+ All Categories
Home > Documents > DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

Date post: 18-Oct-2021
Category:
Upload: others
View: 5 times
Download: 1 times
Share this document with a friend
421
The Pennsylvania State University The Graduate School DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR RADIOFREQUENCY ABLATION AND ADDITIVE MANUFACTURING A Dissertation in Mechanical Engineering by Bradley W. Hanks © 2020 Bradley W. Hanks Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy August 2020
Transcript
Page 1: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

The Pennsylvania State University

The Graduate School

DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR RADIOFREQUENCY

ABLATION AND ADDITIVE MANUFACTURING

A Dissertation in

Mechanical Engineering

by

Bradley W. Hanks

© 2020 Bradley W. Hanks

Submitted in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

August 2020

Page 2: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

ii

The dissertation of Bradley W. Hanks was reviewed and approved by the following:

Mary Frecker

Professor, Mechanical Engineering & Biomedical Engineering

Dissertation Advisor

Chair of Committee

Reuben Kraft

Associate Professor, Mechanical Engineering

Jason Moore

Associate Professor, Mechanical Engineering

Edward Reutzel

Senior Research Associate, Engineering Science and Mechanics, Applied

Research Laboratory

Timothy Simpson

Paul Morrow Professor in Engineering Design and Manufacturing, Mechanical

Engineering, Industrial and Manufacturing Engineering

Karen A. Thole

Professor of Mechanical Engineering

Head of the Department of Mechanical Engineering

Page 3: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

iii

Abstract

Recent advances in the capabilities of model simulation and manufacturing technology

are paving the way for systematic design tools. These tools enable designers to explore organic

and nonconventional designs that were previously not possible to analyze or manufacture.

Through design optimization, these nonconventional designs may be used to maximize part

performance in a variety of applications such as medical, aerospace, and defense industries. One

sector of nonconventional designs that have been relatively unexplored is that of compliant

mechanisms or tailored mechanical properties for metamaterial design.

Compliant mechanisms are structures which gain motion through deflection or

deformation of a material rather than traditional hinges or joints. Metamaterials gain their

properties based their structure as opposed to material composition. Using material deformation

to gain motion allows for generation of unique structures that reduce the wear commonly seen in

hinges, remove the need for lubrication in harsh environments, reduce part count, and improve

dexterity. While these advantages present unique opportunities in a variety of situations, it

remains challenging to design and optimize these nonconventional structures.

In this dissertation two systematic design optimization approaches are described for

compliant structures: a shape matching approach for a deployable surgical tool for radiofrequency

ablation (RFA) of tumors and generation of lattice structures for additive manufacturing (AM).

Radiofrequency ablation is an increasing common minimally invasive treatment option

for abdominal tumors. During the procedure, an electrode is inserted into the tumor and the

surrounding tissue is heated. To effectively destroy a tumor with RFA, it is critical that the shape

of the treatment match the shape of the tumor with a small margin around the periphery. The

shape of the treatment, or ablation zone, is largely dependent on the shape of the electrode.

Page 4: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

iv

To improve treatment for endoscopic RFA of tumors, a novel compliant deployable RFA

electrode is presented along with a systematic design optimization approach for shape matching

the treatment zone and tumor. The optimization approach includes a finite element model of RFA

coupled with a genetic algorithm, an approach that may be applied to tumors throughout the

abdomen and mediastinum. A specific example of the approach is demonstrated through

pancreatic tumor ablation. The results of the approach show treatment efficiency is increased

from 25% to 87% for a 2.5cm spherical tumor. In addition, experimental validation of the finite

element RFA model is reported.

With the recent advances in AM, systematic design optimization approaches are needed

for generating structures which take advantage of the design freedom available. With the added

design freedom, lattice structures, which consist of a repeating unit cell topology patterned

throughout a structure, may be used to reduce weight or increase functionality while maintaining

part performance. For AM, nearly infinite possibilities are available for the unit cell topology that

will populate the lattice structure. However, there is a lack of fundamental understanding of unit

cell topology selection. As a result, lattice structures are used without full consideration for how

the lattice structure changes the properties of the part.

For development of tailored compliance, this dissertation presents a design for AM

(DfAM) systematic design optimization approach for generating manufacturable and customized

unit cell topologies. Using a ground structure topology optimization (GSTO) approach, novel unit

cell topologies may be generated to meet application specific needs. By incorporating a library of

unique optimization objectives, constraints, and penalties, Additive Lattice Topology

Optimization (ALTO) is an application-agnostic approach to unit cell design, demonstrated

through various case studies for unique lattice structure applications.

Two case studies are presented for multi-functional lattices that combine minimization of

strain energy and maximization of thermal conductance. The results show how DfAM

Page 5: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

v

considerations can be included in the optimization to improve manufacturability and are

demonstrated with printed examples of the optimized solution. In addition, these multi-functional

lattices highlight the importance of multi-objective optimization and its ability to generate a

Pareto-optimal set of solutions.

Three other case studies are presented for optimizing structures that match a target

constitutive matrix. The first of these three demonstrates a novel powder removability factor to

generate a unit cell topology that has the same orthogonal effective elastic moduli as a baseline

unit cell topology. The optimized solution shows improved powder removal compared to the

baseline unit cell through mass measurement, CT scan analysis, and tortuosity quantification.

Experimental compression testing of both the optimized and baseline unit cell topologies validate

the predicted results within 6%. The second of these case studies uses optimization to design

three novel unit cell topologies that utilize mechanism-like behavior to reduce stiffness without

reducing weight. This enables lattice structures with a non-uniform stiffness while maintaining a

uniform weight distribution. The third case study demonstrates generation of an orthotropic unit

cell topology, with three times the effective modulus in one direction as compared another.

As an additional part of this work, a compilation of data from analytical models,

experimental characterization, and finite-element analysis of lattice structures has resulted in the

largest public database of mechanical property information for metal lattice structures fabricated

with AM. The Lattice Unit-cell Characterization Interface for Engineers (LUCIE) offers Ashby-

style plots for unit cell topology trends in mechanical properties. It represents a compilation of 69

sources, resulting in 1400 experimental data points, 200+ finite element points, and 45+

analytical models for a total of 18 different unit cell topologies.

The culmination of this work presents research contributions for novel design

optimization approaches for RFA of tumors and DfAM of metal lattice structures through

numerous case studies and experimental validation.

Page 6: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

vi

Table of Contents

List of Figures .......................................................................................................................... ix

List of Tables ........................................................................................................................... xviii

Abbreviations ........................................................................................................................... xx

Nomenclature ........................................................................................................................... xxii

Acknowledgements .................................................................................................................. xxviii

Chapter 1 Introduction ............................................................................................................. 1

1.1 Background and Motivation ....................................................................................... 1 1.1.1 Compliant Mechanisms ................................................................................... 2 1.1.2 Pancreatic Cancer ............................................................................................ 4 1.1.3 Radiofrequency Ablation ................................................................................ 7 1.1.4 Additive Manufacturing .................................................................................. 14 1.1.5 Lattice Structures ............................................................................................. 23 1.1.6 Design Optimization ....................................................................................... 36 1.1.7 Topology Optimization and Additive Manufacturing ..................................... 50

1.2 Research Objectives and Tasks .................................................................................. 53 1.3 Dissertation Contents and Structure ........................................................................... 54

Chapter 2 Optimization and Experimental Validation of an EUS-RFA Electrode .................. 56

2.1 Background and Motivation ....................................................................................... 56 2.2 Deployable Compliant EUS-RFA Electrode Design ................................................. 58 2.3 Systematic Optimization Approach ........................................................................... 61

2.3.1 Ablation Zone Prediction with a Finite Element Model ................................. 61 2.3.2 Optimization using a Genetic Algorithm ........................................................ 65 2.3.3 Results from the Systematic Optimization Approach ..................................... 70 2.3.4 Discussion and Comparison to Standard Straight Electrode ........................... 78 2.3.5 Conclusions from Systematic Optimization Approach ................................... 82

2.4 Testing and Validation of Deployable Electrode ....................................................... 83 2.4.1 Prototype Development ................................................................................... 83 2.4.2 Deployment in a Tissue Phantom .................................................................... 86 2.4.3 Ex-vivo validation of TAM ............................................................................. 93

2.5 Closing Remarks ........................................................................................................ 103

Chapter 3 Unit Cell Generation through Ground Structure Topology Optimization ............... 104

3.1 Background and Motivation ....................................................................................... 104 3.2 Additive Lattice Topology Optimization (ALTO) ..................................................... 107

3.2.1 Ground Structure Model .................................................................................. 108 3.2.2 Optimization Algorithms ................................................................................. 110

Page 7: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

vii

3.2.3 ALTO Objectives, Constraints, and Penalties ................................................. 114 3.2.4 ALTO Validation ............................................................................................ 143

3.3 Closing Remarks ........................................................................................................ 152

Chapter 4 Case Studies on Multi-functional Lattices .............................................................. 153

4.1 Background and Motivation ....................................................................................... 153 4.2 Case Study 1: Generation of a Novel 3D Multi-functional Lattice ............................ 155

4.2.1 Optimization Description ................................................................................ 155 4.2.2 Optimization Results ....................................................................................... 158

4.3 Case Study 2: Generation of a Novel 2D Multi-functional Lattice ............................ 164 4.3.1 Optimization Description ................................................................................ 164 4.3.2 Optimization Results ....................................................................................... 167

4.4 Discussion of Results ................................................................................................. 170 4.5 Closing Remarks ........................................................................................................ 173

Chapter 5 Case Studies on Homogenized Lattice Structures ................................................... 174

5.1 Background and Motivation ....................................................................................... 174 5.2 Case Study 1: Generation of a 3D Unit Cell Topology with Improved Powder

Removability ............................................................................................................ 176 5.2.1 Optimization Description ................................................................................ 177 5.2.2 Optimization Results ....................................................................................... 181 5.2.3 Ground Structure Model Validation ................................................................ 184 5.2.4 Powder Removability ...................................................................................... 193

5.3 Case Study 2: Generation of a 3D Unit Cell Topologies with a Fixed Volume

Fraction .................................................................................................................... 203 5.3.1 Optimization Description ................................................................................ 204 5.3.2 Optimization Results ....................................................................................... 207 5.3.3 Fabrication and Compression Testing ............................................................. 212

5.4 Case Study 3: Generation of a 2D Orthotropic Unit Cell Topology .......................... 214 5.4.1 Optimization Description ................................................................................ 214 5.4.2 Optimization Results ....................................................................................... 218

5.5 Discussion of Results ................................................................................................. 222 5.6 Closing Remarks ........................................................................................................ 228

Chapter 6 LUCIE: Lattice Unit-cell Characterization Interface for Engineers ........................ 229

6.1 Background and Motivation ....................................................................................... 229 6.2 Methods ...................................................................................................................... 231

6.2.1 Literature Compilation .................................................................................... 232 6.2.2 Data Collection ................................................................................................ 238 6.2.3 Assumptions for Comparing Data ................................................................... 240 6.2.4 LUCIE ............................................................................................................. 242

6.3 Results and Case Studies ............................................................................................ 245 6.3.1 Results of Literature Compilation ................................................................... 245 6.3.2 LUCIE Case Study 1 ....................................................................................... 252 6.3.3 LUCIE Case Study 2 ....................................................................................... 253 6.3.4 LUCIE Case Study 3 ....................................................................................... 255

Page 8: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

viii

6.4 Discussion .................................................................................................................. 258 6.5 Closing remarks ......................................................................................................... 264

Chapter 7 Conclusions and Future Work ................................................................................. 266

7.1 Summary of Objectives and Tasks ............................................................................. 266 7.1.1 Objective 1: Develop systematic optimization approach for compliant

deployable radiofrequency ablation electrodes ................................................ 266 7.1.2 Objective 2: Develop three-dimensional ground structure topology

optimization for unit cell DfAM ....................................................................... 267 7.1.3 Objective 3: Explore multi-functional lattices through case study on unit

cell generation for thermal conductance and minimum strain energy .............. 269 7.1.4 Objective 4: Explore homogenized lattices through case study on

generation of unit cells with tailored compliance ............................................. 270 7.1.5 Objective 5: Develop mechanical property database for metal lattice

structures from current literature containing analytical, finite element, and

experimental data ............................................................................................. 271 7.2 Primary Conclusions .................................................................................................. 272 7.3 Research Contributions .............................................................................................. 273

7.3.1 Electrode Optimization for Radiofrequency Ablation .................................... 274 7.3.2 Additive Lattice Topology Optimization ........................................................ 275

7.4 Recommended Future Research Directions ............................................................... 277

References ................................................................................................................................ 283

Appendix A: MATLAB Image Analysis ................................................................................. 316

Appendix B: ALTO 3D MATLAB Files ................................................................................. 319

Page 9: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

ix

List of Figures

Figure 1.1 Compliant surgical forceps reduce the need for complex hinges on a small

scale, lack debris from wear, and do not require lubrication, adapted from [8] ............... 2

Figure 1.2 Examples of multi-material (left) and metal (right) compliant mechanisms

designed for AM, adapted from [13,19] ........................................................................... 4

Figure 1.3 Five year survival rates for the 10 most common cancers [22]. Pancreatic

cancer is the lowest with an average fiver year survival rate of 7%. ............................... 5

Figure 1.4 (Left) Diagram of radiofrequency ablation showing the mismatch that can

occur between a tumor and ablation zone and (Right) thermal-imaging of

radiofrequency ablation showing the ellipsoidal temperature profile around a straight

electrode, adapted from [41] ............................................................................................ 10

Figure 1.5 Basic electrode types and several commercial models of radiofrequency

ablation electrodes described by Mulier et al., images adapted from [51] ....................... 10

Figure 1.6 Metal AM system types include powder bed fusion and wire/powder feed

directed energy deposition, typically using either a laser or electron beam energy

source, adapted from [112] .............................................................................................. 17

Figure 1.7 Example of warping due to thermal contraction..................................................... 23

Figure 1.8 Various forms of cellular solids including a (top) 2D honeycomb, (middle) 3D

open-cell foam, and (bottom) 3D closed-cell foam, adapted from [136] ......................... 25

Figure 1.9 Lattice structures have been proposed for a variety of medical implants with

the intent of improving performance and functionality. Due to the variety of lattice

structures that can be used in such applications, there is a need for further research to

understand how to select or design an appropriate unit cell for the intended

application. Images adapted from A) [179], B) [177], C) [181], D) & E) [194]. ............ 27

Figure 1.10 Unit Cells are often generated based on Boolean operations with primitive

shapes such as these examples, adapted from [195] ........................................................ 28

Figure 1.11 Unit cells may be generated using implicit mathematical equations to

represent a surface. In this particular case, the surface is defined at F(x,y,z) = 0 with

parameters "a" and "b" used to tailor the shape of the surface; adapted from [195]. ....... 29

Figure 1.12 Two common methods for lattice population are (top) uniform in which the

basic lattice is not modified based on the component shape and (bottom) conformal

in which each unit cell may be deformed to follow the surface of the component,

adapted from [224] ........................................................................................................... 30

Figure 1.13 One method of generating a functionally graded lattice is by computing the

stress within the part, plotting the stress in gray scale, and then dithering the image

to generate a density map. Next, small stiff unit cells are placed in dense regions,

Page 10: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

x

with high stress, and large unit cells are placed in less dense regions, adapted from

[229]. ................................................................................................................................ 31

Figure 1.14 Functionally graded lattice structure populated using a relative density map

based on CT imaging, adapted from [178] ....................................................................... 32

Figure 1.15 Lattice generation and sizing method using stress results from a finite

element analysis, adapted from [234] .............................................................................. 33

Figure 1.16 Diagram of genetic algorithm optimization process ............................................. 39

Figure 1.17 Cartoon graphic used to describe maximization of an objective for gradient-

based algorithms, adapted from [254] .............................................................................. 40

Figure 1.18 Minimization of compliance benchmark optimized with the SIMP method.

Varying the power-law penalization factor, p, reduces the amount of gray material in

the solution, clearly defining the boundaries of the topology. Images adapted from

[270]. ................................................................................................................................ 44

Figure 1.19 Boundary and Loading conditions for a compliant inverter as well as the

evolution of the topology during convergence, adapted from [285] ................................ 45

Figure 1.20 A variety of compliant gripper topologies generated by varying optimization

parameters, adapted from [285] ....................................................................................... 46

Figure 1.21 An example comparison of the continuum and ground structure approach for

topology optimization ...................................................................................................... 48

Figure 1.22 Potential tradeoffs exist in the ability to find design that consider the amount

of support material used while simultaneously considering the performance of the

component, adapted from [328] ....................................................................................... 51

Figure 2.1 For straight electrodes, multiple insertions are often required to overlap a

series of ablation zones to destroy tumors. The image above is for an example of a

percutaneous approach, but, for EUS-RFA, the ability to insert multiple electrodes is

limited creating the need for specialized electrodes, adapted from [51]. ......................... 57

Figure 2.2 The electrode is deployed into the tumor through an endoscopic needle, acting

as a sheath, whose tip has been positioned at the periphery of the tumor, adapted

from [339] ........................................................................................................................ 58

Figure 2.3 Two variations of the proposed electrode design are pictured in the stowed

(A,D) and deployed (B,E) configurations. The design parameters for each are shown

(C,F), adapted from [339] ................................................................................................ 60

Figure 2.4 General setup for the TAM. Due to the symmetry of the electrode, the TAM

was reduced to quarter symmetry to decrease computation time. The grounding pads

regions are the top and bottom quarter circles as well as the exterior wall of the

cylinder (not shown) , adapted from [339]. ...................................................................... 64

Page 11: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

xi

Figure 2.5 Graphical description of the objective functions used for optimization, adapted

from [339] ........................................................................................................................ 68

Figure 2.6 Each design tested during the optimization is plotted in this figure according to

its corresponding objective function values, adapted from [339] .................................... 74

Figure 2.7 (Left) The side view of the temperature profile for an optimal electrode

solution for Case 1. The ablation zone is marked by the white isothermal contour

(60°C). (Right) The ablation zone surface (green) is shown compared to the target

shape (black outline) for the top and side views, adapted from [339]. ............................ 75

Figure 2.8 Each Case 2 design simulated is plotted above according to its corresponding

objective function values, adapted from [339] ................................................................. 76

Figure 2.9 (Left) The side view of the temperature profile for an optimal electrode

solution for Case 2. The ablation zone is marked by the white isothermal contour

(60°C). (Right) The ablation zone surface (green) is shown compared to the target

shape (black outline) for the top and side views, adapted from [339]. ............................ 77

Figure 2.10 Top and side views of the ablation zones for the straight electrode and

optimal solutions of Case 1 and Case 2. The green surface represents the 60°C

isothermal surface and the black circle represents the target zone, adapted from

[339]. ................................................................................................................................ 80

Figure 2.11 L-PBF 2X scale electrode prototype .................................................................... 84

Figure 2.12 Near 1X scale L-PBF electrode prototype ............................................................ 84

Figure 2.13 4-tine laser micromachined prototype .................................................................. 85

Figure 2.14 8-tine and 12-tine prototypes based on optimal solution configurations that

used for ex-vivo experimentation..................................................................................... 86

Figure 2.15 Insertion force and deployment experimental setup for early prototypes of the

EUS-RFA electrode ......................................................................................................... 87

Figure 2.16 (Left) The spread or amount of deployment is shown based on the insertion

speed. (Right) The insertion force profile is separated into three segments showing

the friction sliding through the sheath, the penetration of the central needle, and then

the deployment of the outer tines. Both graphs are for the L-PBF electrode. .................. 89

Figure 2.17 (Left) The spread or amount of deployment is shown based on the insertion

speed. (Right) The insertion force profile is separated into three segments showing

the friction sliding through the sheath, the penetration of the central needle, and then

the deployment of the outer tines. Both graphs are for the LMM electrode. ................... 90

Figure 2.18 Geometry details for 8-tine deployable electrode prototype for TAM

validation, adapted from [355] ......................................................................................... 94

Page 12: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

xii

Figure 2.19 Schematic and photo of the experimental set-up for thermochromic tissue

phantom testing, adapted from [355] ............................................................................... 96

Figure 2.20: Tissue phantom and TAM ablation zone and analyzed images for the 15W

power level, adapted from [355]. (Scale 15.4 px/mm) ..................................................... 99

Figure 2.21: Tissue phantom and TAM ablation zone and analyzed images for the 20W

power level, adapted from [355]. (Scale 17.4 px/mm) ..................................................... 99

Figure 2.22: Comparison of the ablation zone size and shape by the major and minor axis

of an ellipse fitted to each ablation zone with image analysis; see examples in Figure

2.20 and Figure 2.21, adapted from [355] ........................................................................ 100

Figure 3.1 Overview of Additive Lattice Topology Optimization (ALTO) to generate

optimized unit cell topologies for lattices fabricated with AM ........................................ 108

Figure 3.2 Example of traditional thermal circuits and analogy to the ground structure

formulation for calculating the unit cell thermal conductance ......................................... 118

Figure 3.3 Schematic for three test strain cases to an RVE for calculation of the

homogenized constitutive matrix, adapted from [298] .................................................... 121

Figure 3.4 Schematic showing how powder can become trapped in internal cavities for

powder-based AM processes............................................................................................ 125

Figure 3.5 Ground structure-based pore size measurement for the PRF. On the left part of

the image, an example of the element-to-element distance measurement is made

midpoint to midpoint and then the radius of each element is subtracted to obtain the

pore size for the pore distance matrix. A few examples of some of the pore

measurements are shown on the right part of the image. In the pore matrix, each

element is measured relative to every other element. ...................................................... 126

Figure 3.6 For lattice structures, horizontal elements are allowed to span between two

vertical supports because of the generally small length-scale. However, there is an

intermediate range for which the build angle is too low to accurately build. .................. 130

Figure 3.7 Horizontal lattice members are shown in two lattice examples, 3x3x3 arrays of

octet-truss unit cells (4 mm, 7 mm) and cubic unit cells (4 mm, 6 mm). These

lattices were printed in Ti64 on a 3D Systems ProX DMP 320. Image adapted from

[361]. ................................................................................................................................ 131

Figure 3.8 Basic 2D example for determining the supported nodes for the unsupported‒N

penalty. This example comes from row three of Table 3.2 which passes the

unsupported‒N penalty but would fail the unsupported‒S penalty. The unsupported‒

S penalty follows a similar process but the node-neighbor pairs would not be

considered meaning that node 4 is not self-supported, but instead supported by a

neighboring unit cell. Horizontal elements that bridge supported nodes are

considered feasible. .......................................................................................................... 140

Page 13: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

xiii

Figure 3.9 Graph of penalty function that reduces complexity by encouraging elements

with a cross-sectional area larger than the relevant cross-sectional area limit. ................ 142

Figure 3.10 Schematics for the MBB beam benchmark problem and the initial ground

structure used to represent the design space..................................................................... 145

Figure 3.11 Schematics for the cantilever beam benchmark problem and the initial

ground structure used to represent the design space. ....................................................... 145

Figure 3.12 MBB beam comparison with results reported in literature. The different

solutions are shown for a fixed volume and changing the minimum feature size

limit. The line width on the bottom images is correlated to the cross-sectional area

from the GSTO solutions. Top row of images adapted from [273]. ................................ 146

Figure 3.13 Cantilever benchmark comparison with literature results. Various solutions

are shown for a fixed allowable volume and changing the minimum feature size

limit. The line width on the bottom images is correlated to the cross-sectional area

from the GSTO solutions. Top row of images adapted from [273]. ................................ 147

Figure 3.14 Schematic for thermal conductance benchmark from literature. The

continuum topology optimization approach (left) is shown compared to the GSTO

approach (right). ............................................................................................................... 149

Figure 3.15 Thermal conductance benchmark results from literature (left, middle) and

from the GSTO in ALTO show good correlation, validating the thermal conductance

formulation and optimization objective. Left and middle images adapted from [278]

and [281], respectively. .................................................................................................... 150

Figure 4.1 Case Study 1: 3D multi-functional lattice example problem, simultaneously

optimizing for (A) minimization of strain energy and (B) maximizing the thermal

conductance ...................................................................................................................... 157

Figure 4.2 Solutions to the optimization problem from Case Study 1. Design A and

Design B are the solutions without and with the DfAM penalty, respectively. ............... 159

Figure 4.3 Design A has several long horizontal overhangs reducing the

manufacturability of the unit cell. Incorporating a DfAM penalty into the

optimization, the algorithm discovered Design B which is fully self-supported. The

images depict an XZ plane through the center of the unit cell topology for improved

visibility of some of the unsupported regions. ................................................................. 159

Figure 4.4 Graphs of the solutions space for an instance of the Case Study 1

Optimization. Each point is a Pareto-optimal solution for (A) modified thermal

conductance versus strain energy objectives and (B) modified thermal conductance

versus volume fraction objectives. The red point in each graph represents Design B

shown in Figure 4.2 .......................................................................................................... 162

Figure 4.5 Printed examples of the optimized solution from the 3D multi-functional

lattice case study. The examples were printed as a single unit cell (scaled up 5x)

with FFF and in a 2x2x2 array (scaled 2x) with L-PBF. By incorporating DfAM

Page 14: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

xiv

considerations into the optimization, the optimized solution was print-ready, not

requiring any design interpretation. ................................................................................. 164

Figure 4.6 Case Study 2: (A) Minimize strain energy and (B) Maximize thermal

conductance. Blue dashed lines denote symmetry planes. Images adapted from

[361]. ................................................................................................................................ 166

Figure 4.7 Example convergence plot for the solution to the optimization problem with

equal weighting factors for both objectives. The result shows an increasing objective

function score until the constraints are met and then gradually decreasing until

convergence. .................................................................................................................... 168

Figure 4.8: Case Study 2 results showing the tradeoffs between minimization of strain

energy and maximization of thermal conductance. 11 different solutions are shown

beginning for incrementing the weight values by 0.1 as the solutions move

clockwise. Image adapted from [361]. ............................................................................. 168

Figure 4.9: Comparison of the objective function values for minimization of strain energy

and maximization of thermal conductivity (minimization of its’ reciprocal plotted

here). This plot shows the solution space, demonstrating the tradeoffs between

objectives. On the plot, all 101 solutions are shown, with the grey dots being non-

dominated solutions on the Pareto-front. Image adapted from [361]. .............................. 169

Figure 5.1 The octet-truss (A) is the baseline unit cell topology for matching the

constitutive matrix, colors have been added to improve visualization of 3D topology.

The initial ground structure is shown in (B), where each line represents a truss

element. With symmetry, there were a total of 120 distinct design variables. ................. 181

Figure 5.2 The two unit cells output from the optimization are shown on either side of the

octet-truss. The top images show a single unit cell and the bottom images are side

views of a 3x3 patterned array. As a result of the PRF as an objective, both the

octahedral and opti-octet open up the porous space to improve powder removal and

flowability. ....................................................................................................................... 182

Figure 5.3 Complete build plate with octet-truss and opti-octet lattice structures among a

few other lattice types. In the picture, one of the five samples of each lattice type are

labeled. The interlocked and diamond lattice are discussed as part of Case Study 2 in

Section 5.3. ....................................................................................................................... 185

Figure 5.4 Examples of the octet-truss (left) and opti-octet (right) crushed after

compression loading. Prior to compression, the samples were 40 x 40 mm cubes or a

unit cell size of 5mm as noted by the scale bars. ............................................................. 186

Figure 5.5 Average stress strain curves for the octet-truss and opti-octet truss. The solid

portion of the curves represents the region where the effective modulus was

measured, 10-50% of the peak stress. The grey band behind each curve represents

the average plus and minus the standard deviation of the experimental data. ................. 187

Figure 5.6 AM-based corrections applied to ground structure prediction. As a result of

either or both corrections, the cross-section of the strut is effectively reduced. .............. 190

Page 15: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

xv

Figure 5.7 Schematic of a compression sample where the color gradient represents the

impact of edge effects from the friction with compression platens. Near the

compression platens a unit cell experiences “constrained” boundary conditions as

compared to “relaxed” boundary conditions for a unit cell near the middle of the

sample. ............................................................................................................................. 192

Figure 5.8 As-printed lineup of octet-truss powder removal specimens. The largest

specimen is a 5x5x5 array of 7 mm unit cells, approximately 35 mm in width. The

smallest specimen is a 5x5x5 array of 1.5 mm unit cells, approximately 7.5 mm in

width. ............................................................................................................................... 195

Figure 5.9 Comparison of the relative error between the as-designed and as-printed mass

for the powder removal specimens. In (B) three regions are predicted based on the

relative error (1) no powder removed, (2) partial removal, and (3) all powder

removed, shown with the red ellipses overlaid. ............................................................... 197

Figure 5.10 Qualitative CT scan comparison of a central slice for the octet-truss and opti-

octet unit cell sizes. The dashed line squares represent a single unit cell within each

image. Comparing the 4.5 mm unit cell sizes against the computer generated as-

designed cross-section, the octet-truss shows more trapped powder or partially

sintered regions as compared to the opti-octet, validating that the opti-octet

demonstrates improved powder removability. ................................................................. 200

Figure 5.11 Case Study 2 is a multi-objective optimization for creating a lattice structure

that mimics the homogenized constitutive matrix of a Diamond unit cell (A) and (B).

The unit cell graph is shown in (C) and (D). To ensure similar interconnectivity to

the Diamond unit cell, the full ground structure is reduced to the initial configuration

shown in (E) and (F). ....................................................................................................... 207

Figure 5.12 Case Study 2 resulted in three unit cells which match the scaled target

constitutive matrix of the diamond unit cell. In each solution, a fundamental

tetrahedra (grey) shape was found with decreasing diameter. For the middle and top

solutions, several non-load bearing elements (blue) form a mechanism around the

tetrahedra shape, based on the pin-joint assumption, to simultaneously match the

volume fraction and homogenization targets. .................................................................. 208

Figure 5.13 The interlocked lattice is a unique patterning of the tetrahedra base that

results in a scaled and interlocked set of two diamond lattice. In the final patterned

interlocked lattice, the unit cells only connect to diagonally neighboring unit cells. ...... 210

Figure 5.14 The Interlocked lattice is a set of two interconnected lattices. The upper

lattices were printed in Alumide® on an EOS P110. The lower left image shows the

interlocking lattices, distinguished by the blue and grey colors. ..................................... 211

Figure 5.15 Average stress strain curves for the diamond and interlocked lattices. The

solid portion of the curves represents the region where the effective modulus was

measured, 7.5-60% of the peak stress. The grey band behind each curve represents

the average plus and minus the standard deviation of the experimental data. ................. 213

Page 16: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

xvi

Figure 5.16 The purpose of Case Study 3 is to generate a unit cell topology that has an

orthotropic constitutive matrix, specifically that it is three times stiffer in the X-

direction than in the Y-direction. ..................................................................................... 214

Figure 5.17 Graph of the DfAM penalty function that penalizes elements near the

relevant cross-sectional area limit and reduces non-critical elements. Image adapted

from [361]. ....................................................................................................................... 217

Figure 5.18 Lowest optimization scores for the constitutive matching objective function.

Each image represents the best solution of 20 initial guesses for the particular

combination of penalty factor level and relevant diameter limit. Image adapted from

[361]. ................................................................................................................................ 219

Figure 5.19 Optimized unit cell topology with the area of the elements connected to the

corners set to the lower limit. Compression in X and Y is also shown, demonstrating

the target constitutive matrix. Total error in the constitutive matrix was 3x10-6 %.

Image adapted from [361]. ............................................................................................... 221

Figure 5.20: Example of a lattice structure patterned from the unit cell topology shown in

Figure 5.19. This unit cell topology was generated using a penalty function level of

one third and the relevant diameter limit set to 0.5 mm. Image adapted from [361]. ...... 221

Figure 5.21 Without DfAM restrictions and penalties to ensure manufacturability, there

were a variety of pitfalls that appear in optimization solutions. The DfAM

considerations help to reduce the need for design interpretation and prepare more

print-ready solutions. ....................................................................................................... 226

Figure 6.1 Graphic summary of the methods section including how sources were found

and refined, data collection and assumptions made, and finally incorporation into

LUCIE. ............................................................................................................................. 232

Figure 6.2 Schematic showing the inputs and outputs for LUCIE .......................................... 244

Figure 6.3 Typical stress-strain curves for stretch-dominated or bend-dominated

structures, labeled with the relevant mechanical property terminology for the data

compiled in this work ....................................................................................................... 249

Figure 6.4 Relative frequency of uncertainty in experimental data points. Uncertainty was

normalized by the nominal value to obtain the relative error. 75% of the reported

uncertainty values had a relative error of less than 10%. ................................................. 252

Figure 6.5 LUCIE Case Study 1 output comparing analytical, experimental, and FEA

data for the BCC, Diamond, and Truncated Cube unit cell topologies. From the

output you can visualize differences in analytical models for a single unit cell

topology, such as the Diamond, as well as gain an understanding of the normalized

effective modulus for different unit cell topologies. ........................................................ 253

Figure 6.6 LUCIE Case Study 2 output shows a comparison of 7 different unit cell

topologies and the variability that exists in experimental data. The error bars around

a single point represent a small “bin” of data. The point itself represents the average

Page 17: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

xvii

value within the bin, while the error bars denote the maximum and minimum

reported values for each property. .................................................................................... 254

Figure 6.7 LUCIE Case Study 3 output from LUCIE for comparison of the lattice

structure properties of different unit cell topologies. The red box with a dashed line

highlights the region of interest (EC 0.02-0.03 and Volume Fraction 0.2-0.25) for

determining appropriate unit cell selection. ..................................................................... 256

Figure 6.8 LUCIE Case Study 3 plot comparison of normalized effective modulus and

normalized yield stress for comparison of the six unit cell topologies that were down

selected from Figure 6.7. The thumbnails of the unit cell topologies have been added

to the plot; from top to bottom they are Cubic, TPMS Gyroid,

Rhombicuboctahedron, Truncated Cuboctahedron, BCC. No yield stress data was

available for the Octet-truss at this volume fraction range. ............................................. 257

Figure 6.9 Number of sources found for each unit cell topology. The results show the

amount of lattice structure data available is limited for many common unit cell

topologies. ........................................................................................................................ 262

Page 18: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

xviii

List of Tables

Table 1.1 A wide range of opportunistic and restrictive DfAM topics exist for which

there is ongoing research .................................................................................................. 19

Table 2.1 Initial and boundary conditions for the TAM .......................................................... 64

Table 2.2 Material properties for TAM ................................................................................... 71

Table 2.3 Case 1 and Case 2 optimization and electrode design parameters ........................... 72

Table 2.4 Electrode designs selected for comparison from the Case 1 and Case 2 Pareto-

optimal solution sets ......................................................................................................... 77

Table 2.5 Quantitative comparison of a straight electrode and optimal Case 1 and Case 2

solutions ........................................................................................................................... 79

Table 2.6: Tissue phantom material properties [357] .............................................................. 95

Table 3.1 Summary of ALTO objective, constraint, and penalty functions for unit cell

generation and their applicability to DfAM ..................................................................... 115

Table 3.2 Examples of 2D unit cells for comparing pass/fail criteria of different penalty

functions. .......................................................................................................................... 138

Table 3.3 Comparison of homogenization benchmark problems and ground structure

calculation. For these benchmark problems, the ground structure was an exact match

to the reported constitutive matrix, images adapted from [297,298] ............................... 151

Table 4.1 Case Study 1: 3D Multi-functional lattice optimization summary .......................... 155

Table 4.2 Comparison of objective function values for two designs from Case Study 1.

Design A is not manufacturable due to long unsupported overhangs in the topology. .... 160

Table 4.3 List of Pareto-optimal solutions with the modified thermal conductance

objective less than -0.10, strain energy objective less than 95, and the volume

fraction objective sorted smallest to largest. According to the requirements of the

designer, the optimal solution would be Design B shown in the first row in bold

typeface. ........................................................................................................................... 163

Table 4.4 Case Study 2: 2D Multi-functional lattice optimization summary .......................... 165

Table 5.1 Case Study 1: Improved powder removal lattice optimization summary ................ 178

Table 5.2 Summary of the objective function values for the optimization results. The

octet-truss results are shown for individual evaluation of the unit cell, separate from

the optimization. ............................................................................................................... 182

Page 19: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

xix

Table 5.3 Experimental validation results for the effective modulus and comparison to the

ground structure model prediction in ALTO and AM-based corrections. For AM

corrections, IO = infill-only correction and SC = stair case correction. The error

relative to the experimental value is noted in the parenthesis. ......................................... 187

Table 5.4 Effective modulus comparison between experimental data and ALTO

predictions of “constrained” and “relaxed” boundary conditions. The error relative to

the experimental value is noted in the parenthesis. .......................................................... 193

Table 5.5 Case Study 2: 3D homogenized unit cells with different stiffnesses

optimization summary ...................................................................................................... 205

Table 5.6 Case Study 3: 2D orthotropic unit cell generation optimization summary .............. 215

Table 5.7 Lattice structures demonstrate a substantial difference in mechanical properties

based on the number of layers. As the number of layers increase, the effective

modulus decreases, data from [207]. ................................................................................ 223

Table 6.1 For sources without bulk mechanical properties of the as-built material,

reported within the source, these generic AM mechanical property values were used

to normalize data to create a material independent comparison of unit cell topology. .... 234

Table 6.2 Summarized list of all sources for data compiled in LUCIE. In this table the

unit cell topology, properties recorded, and data type available are sorted by source. .... 236

Table 6.3 For each unit cell topology included in LUCIE, there are 6 different properties

that can be compared. The data type available for each unit cell topology are denoted

by analytical (A), experimental (E), or FEA (F). A dash (-) signifies that no

reference was found that provided this information for any of the three data types. ....... 246

Table 6.4 Name and associated picture for each lattice structure included in LUCIE ............ 247

Page 20: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

xx

Abbreviations

2D Two-dimensional

3D Three-dimensional

A Analytical data type for LUCIE

AM Additive manufacturing

AMI Actuated Medical Incorporated

BCC Body-centered cubic

BS Lattice structure buckling stress normalized by the solid yield stress

CAD Computer-aided design

CM Compliant mechanism

CT Computed tomography

DED Directed energy deposition

DfAM Design for additive manufacturing

E Experimental data type for LUCIE

EC Lattice structure compressive effective elastic modulus normalized by the solid

material modulus

EUS-RFA Endoscopic ultrasound-guided radiofrequency ablation

F Finite element analysis data type for LUCIE

FCC Face-centered cubic

FEA Finite element analysis

FFF Fused filament fabrication

GSTO Ground structure topology optimization

L-PBF Laser powder bed fusion

LMM Laser micromachining

LUCIE Lattice unit-cell characterization interface for engineering

MRI Magnetic resonance imaging

NSGA-II Non-dominated sorting genetic algorithm

PBF Powder bed fusion

PR Lattice structure Poisson’s ratio

Page 21: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

xxi

PS Lattice structure plateau stress normalized by the solid yield stress

PVC Polyvinyl chloride

PRF Powder removability factor

RFA Radiofrequency ablation

RVE Representative volume element

SIMP Simplified isotropic material with penalization

STL Stereolithograhy file type, a common model file for 3D-printing

TAM Thermal ablation model

TPMS Triply periodic minimal surfaces

US Ultrasound

YS Lattice structure yield stress normalized by the solid yield stress

Page 22: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

xxii

Nomenclature

𝐴𝑒 Cross-sectional area of each element

𝐴𝑙𝑜𝑤𝑒𝑟 Lower limit of cross-sectional area

𝐴𝑟𝑒𝑙 Relevant cross-sectional area limit

𝐴𝑢𝑝𝑝𝑒𝑟 Upper limit of cross-sectional area

𝑨 Vector of element cross-sectional areas

𝑩 Connectivity vector

𝑐𝑖 Weight values for combining optimization objectives

𝐶 Specific heat of pancreatic tissue

𝐶𝑏 Specific heat of blood

𝐶𝑙𝑖 Constraint value for relative center and outer tine length

𝑑𝑒 Stowed electrode outer diameter

𝑑𝑙𝑜𝑤𝑒𝑟 Lower diameter limit

𝑑𝑢 Distance to the utopia point

𝑑𝑢𝑝𝑝𝑒𝑟 Upper diameter limit

𝑑𝑟𝑒𝑙 Relevant diameter limit or minimum feature size

𝒅0𝑒𝑘 Induced displacement vector for element “e”, subject to test strain case “k”

𝒅𝑒𝑘 Applied displacement vector for element “e”, subject to test strain case “k”

𝐷𝑖𝑗𝑒𝑒 Distance from midpoint to midpoint between element i and element j

�̃�𝑒𝑒 Element-to-element distance matrix

𝐸 Modulus of elasticity

�̃� Material constitutive matrix

�̃�𝐻 Homogenized constitutive matrix

�̃�𝐷𝐻 Diamond unit cell homogenized constitutive matrix

�̃� Overlap penalty matrix

�̃�∗ Target constitutive matrix

F Ground structure force vector

𝑓 Overall objective function for gradient-based optimization algorithm

Page 23: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

xxiii

𝑓𝑖 Percentage greater or smaller than 𝑡𝑖 for a grid point

𝑓𝑜𝑣𝑒𝑟 Objective function representing ablation zone outside the target region

𝑓𝑢𝑛𝑑𝑒𝑟 Objective function representing ablation zone inside the target region

ℎ𝑏 Height of insulated shaft connected to the electrode

ℎ𝑡 Height of biological tissue during simulation

𝑱 Current Density

𝑱𝑒 Externally generated current density

𝑘 Thermal conductivity

�̃�𝑒 Element stiffness matrix for element “e” (4x4 for 2D, 6x6 for 3D)

�̃� Ground structure global stiffness matrix

𝑙𝑏 Exposed electrode base length

𝑙𝑐 Electrode center tine length

𝐿𝑒 Ground structure element length

𝑙𝑜 Length of outer tines

𝑙𝑐𝑙𝑜𝑤𝑒𝑟 Lower bound of center tine length

𝑙𝑐𝑢𝑝𝑝𝑒𝑟

Upper bound of center tine length

𝑙𝑜𝑒𝑣𝑒𝑛 Length of even numbered electrode tines for Case 2

𝑙𝑜𝑖 Electrode outer tine length, 𝑖 = 1,… , 𝑛𝑡

𝑙𝑜𝑙𝑜𝑤𝑒𝑟 Lower bound of outer tine length

𝑙𝑜𝑜𝑑𝑑 Length of odd numbered electrode tines for Case 2

𝑙𝑜𝑢𝑝𝑝𝑒𝑟

Upper bound of center tine length

𝐿𝑒 Ground structure element length

𝐿𝑙𝑜𝑤𝑒𝑟 Lower limit of ground structure element length

𝐿𝑢𝑝𝑝𝑒𝑟 Upper limit of ground structure element length

𝑳 Vector of element lengths

𝑚𝑖 Minimizes (𝑚𝑖 = 1) or maximizes (𝑚𝑖 = −1) an objective function

𝑀 Total number of elements in the ground structure

𝑛𝑅𝐸 Number of active elements in the ground structure (𝐴𝑒 > 𝐴𝑟𝑒𝑙)

Page 24: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

xxiv

𝑛𝑔𝑝 Number of grid points for computing the objective functions

𝑛𝑜𝑏𝑗 Number of objective functions in the optimization

𝑛𝑜𝑣𝑒𝑟 Number of grid point locations outside the target region

𝑛𝑝 Number of penalty functions in the optimization

𝑛𝑠𝑡𝑎𝑙𝑙 Number of stall generations for determining convergence

𝑛𝑡 Number of electrode outer tines

𝑛𝑢𝑛𝑑𝑒𝑟 Number of grid point locations inside the target region

𝒏 Surface normal vector

𝑁𝑡 Total number of nodes in the ground structure

𝑁1, 𝑁2 Node (or set of nodes) for thermal conduction

𝑁𝑑𝑜𝑓 Number of degrees of freedom

𝑶 Null vector

�̃� Null matrix

𝑝∗ Pore threshold for powder removability factor

𝑝𝐶 Connectivity penalty

𝑝𝑖 Individual penalty scores

𝑝𝑂 Overlap penalty

𝑝𝑠𝑒 Penalty for each active element (gradient-based only)

𝑝𝑈𝑁 Unsupported–N penalty

𝑝𝑈𝑆 Unsupported–S penalty

𝑃 Combined penalty value

�̃� Pore distance matrix for powder removability

𝒒 Heat flux

𝑞𝑒𝑓𝑓 Effective heat flux from 𝑁1 to 𝑁2 for thermal conduction

�̃�𝒆 Element mutual energy matrix for element “e”

𝑄𝑒𝑥𝑡 Heat source from spatial heating

𝑄𝑗 Current source

𝑄𝑚𝑒𝑡 Metabolism heat source

Page 25: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

xxv

𝑟𝑏 Radius of insulated shaft connected to the electrode

𝑟𝑖 Radial distance from the center of the ablation zone to a grid point

𝑟𝑜 Radius of curvature of outer tines

𝑟𝑡 Radius of biological tissue during simulation

𝑟𝑖𝑒 Radius of element i in ground structure

𝑟𝑜𝑒𝑣𝑒𝑛 Radius of curvature for even numbered electrode tines for Case 2

𝑟𝑜𝑖 Electrode outer tine radius of curvature, 𝑖 = 1,… , 𝑛𝑡

𝑟𝑜𝑙𝑜𝑤𝑒𝑟 Lower bound of outer tine radius of curvature

𝑟𝑜𝑜𝑑𝑑 Radius of curvature for odd numbered electrode tines for Case 2

𝑟𝑜𝑢𝑝𝑝𝑒𝑟

Upper bound of outer tine radius of curvature

𝑅𝑒 Element thermal resistance

𝑠𝑖 Penalty function scalar to determine impact of penalty

𝑺 Unsupported elements vector

𝑺𝑁 Unsupported elements vector, augmented by neighbors

𝑆𝐸 Strain energy

𝑇 Temperature

�̅� Average temperature at heat input nodes for modified thermal conduction

𝑇1, 𝑇2 Temperature assigned to corresponding 𝑁1, 𝑁2 node(s) for thermal conduction

𝑇𝑏 Blood temperature

𝑡𝑖 Target radial distance from the center of the ablation zone to a grid point

𝑇∞ Ambient tissue temperature

𝒖 Ground structure nodal displacement vector

𝒗 Velocity vector

𝑉 Voltage potential

𝑉𝑢𝑐 Unit cell volume

𝑉𝑓 Volume fraction

𝑉𝑓∗ Target volume fraction

𝑉𝑚𝑎𝑥 Total volume of all ground structure elements at 𝐴𝑢𝑝𝑝𝑒𝑟

Page 26: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

xxvi

𝑉0 Voltage applied to the electrode

𝑤𝑖 Individual objective score

𝑤𝑅𝐸 Relevant elements objective

𝑤𝐻 Homogenization objective function

𝑤𝑀𝑇𝐶 Modified thermal conduction objective function

𝑤𝑀𝑇𝐶̅̅ ̅̅ ̅̅ ̅ Normalized modified thermal conduction objective function

𝑤𝑃𝑅𝐹 Powder removability factor objective function

𝑤𝑆𝐸̅̅̅̅ Normalized strain energy objective function

𝑤𝑆𝐸𝑚𝑖𝑛 Strain energy for all element areas set to 𝐴𝑢𝑝𝑝𝑒𝑟

𝑤𝑆𝐸𝑚𝑎𝑥 Strain energy for all element areas set to 𝐴𝑙𝑜𝑤𝑒𝑟

𝑤𝑇𝐶 Thermal conduction objective function

𝑤𝑇𝐶̅̅̅̅ Normalized thermal conduction objective function

𝑤𝑇𝐶𝑚𝑖𝑛 Thermal conduction for all element areas set to 𝐴𝑙𝑜𝑤𝑒𝑟

𝑤𝑇𝐶𝑚𝑎𝑥 Thermal conduction for all element areas set to 𝐴𝑢𝑝𝑝𝑒𝑟

𝑤𝑉 Volume objective function

𝑤𝑉𝐹 Volume fraction objective function

𝑊𝑖 Penalized objective function score

𝑊 Combined objective function value

𝑌 Characteristic length of the representative volume element

𝛼 Ratio of 𝐸𝑥𝑥/𝐸𝑦𝑦 for 2D constitutive matrix

𝛼1, 𝛼2, 𝛼3, 𝛼4 Parameters used to define the penalty function for gradient-based algorithms

𝛽1, 𝛽2, 𝛽3 Property-volume fraction constants for power law regression model

𝛾 Penalty impact level

𝜖 Threshold or tolerance on the objective function values for Borg

𝜂 Tolerance for Pareto front distribution change during 𝑛𝑠𝑡𝑎𝑙𝑙 generations

𝜈 Poisson’s ratio

𝜙 Spherical coordinates angle from Z axis to direction vector

𝜌 Density of pancreatic tissue

Page 27: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

xxvii

𝜌𝑏 Density of blood

𝜎 Electrical conductivity of pancreatic tissue

𝜃 Spherical coordinates angle from X axis in XY plane

𝜃𝑒 Element build angle measured from the build plate

𝜃𝑙𝑜𝑤𝑒𝑟 Lower limit for build angle constraint

𝜃𝑢𝑝𝑝𝑒𝑟 Upper limit for build angle constraint

𝜓 Mechanical property for LUCIE power law regression

𝜔𝑏 Perfusion rate of blood

Page 28: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

xxviii

Acknowledgements

The completion of this dissertation was a process that started long before my first day at

the Pennsylvania State University and would not have been possible without so much help along

the way. First, I would like to acknowledge the mentorship of my adviser Dr. Mary Frecker. Her

continued support and guidance throughout my MS and PhD programs have been invaluable,

without which, I would not have been pushed to become a better researcher, scholar, and person.

Dr. Frecker has opened my eyes to design optimization and I’ll never be able to look at another

problem again without wondering about the tradeoffs in the design and how it could be

optimized. In addition, I am grateful for the feedback and recommendations of my committee

members, Dr. Kraft, Dr. Moore, Dr. Reutzel, and Dr. Simpson which have helped shape my

research.

I would like to acknowledge my undergraduate institution, Brigham Young University.

The education and opportunities there strengthened my drive for engineering and laid the

groundwork for my education. Working with Professor Brian Jensen and various graduate and

undergraduate students in the BioMEMS group solidified my love of research, ultimately leading

to my decision to pursue a PhD.

During my time at Penn State, I have had so much help from many different groups and

individuals such as the EDOG Lab with all of its students, past and present. I’m grateful for the

support of Dr. Simpson, through attending the additive manufacturing research group meetings

and as part of the Additive Manufacturing and Design (AMD) program. It was my first parts

printed in the EDOG lab and later at CIMP-3D that spurred my fascination for additive

manufacturing which later led me a major portion of my research and contributions to the

scientific community. I appreciate the support of Dr. Moyer who provided invaluable insight for

the development of radiofrequency ablation electrodes. I’m grateful for support from Dr. Ounaies

Page 29: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

xxix

and Dr. vonLockette and their labs in the beginnings of my research. I would like to thank Joseph

Calogero whose support in coding and optimization saved me countless hours of learning to

program and debugging. I would like to thank the undergraduate students I mentored and

collaborated with during my research including multiple Capstone teams and more directly,

Katherine Reichert, Fariha Azhar, and Joseph Berthel. I have also had the pleasure of working

with various collaborators outside of Penn State which have made this research possible including

Kevin Snook, Ryan Clement, and Jenna Greaser from Actuated Medical Inc. and Greg Hayes and

the entire Additive Minds Team from EOS North America.

Numerous sources have also supported this work financially including the National

Science Foundation (NSF) EFRI grant #1240459, Air Force Office of Scientific Research

(AFOSR), National Institutes of Health (NIH) award #CA225169, Actuated Medical Inc., NSF

Center for Healthcare Organization (CHOT), CIMP-3D, EOS Additive Minds, Industry 4.0 Seed

Grant, Kulakowski Travel Grant, Reiss Fellowship, Marley Fellowship, and the University

Graduate Fellowship from the Pennsylvania State University. Any opinions, findings, and

conclusions or recommendations expressed in this dissertation are those of the author and do not

necessarily reflect the views of the funding agencies.

Finally, I acknowledge the never-ending support of my family. I’m grateful to my wife,

Brooke, for her encouragement, understanding, and for always listening to my successes and

challenges, even when I didn’t do a great job of explaining them. My parents, Byron and Patti

Hanks for always encouraging me to learn all I can and their support and confidence in me to

accomplish what I set my mind to. I’m thankful for my kids that help me take a break from

thinking of research 24/7 and my extended family for their endless encouragement. Finally, I’m

grateful for God, who I feel has blessed and inspired me throughout my life.

Page 30: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

1

Chapter 1 Introduction

1.1 Background and Motivation

Recent advances in the capabilities of model simulation and manufacturing technology

are paving the way for systematic design tools. These tools enable designers to explore organic

and nonconventional designs that were previously not possible to analyze or manufacture.

Through design optimization, these nonconventional designs may be used to maximize part

performance in a variety of applications such as medical, aerospace, and defense industries. One

sector of nonconventional designs that has been relatively unexplored is that of compliant

mechanisms or components with tailored properties specific to the application.

The goal of this research is to develop systematic design optimization approaches for

compliant structures. In this dissertation, two systematic design optimization approaches are

described: a shape matching approach for a compliant deployable surgical tool for radiofrequency

ablation (RFA) of tumors and the development of specialized lattice structures for additive

manufacturing (AM). The five objectives of this research are as follows:

1. Develop systematic optimization approach for compliant deployable radiofrequency

ablation electrodes

2. Develop three-dimensional ground structure topology optimization for unit cell DfAM

3. Explore multi-functional lattices through case study on unit cell generation for thermal

conductance and minimum strain energy

4. Explore homogenized lattices through case study on generation of unit cells with tailored

compliance

5. Develop mechanical property database for metal lattice structures from current literature

containing analytical, finite element, and experimental data

Page 31: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

2

This chapter includes relevant literature for the optimization approaches discussed in the

dissertation. There are six major sections in the chapter: compliant mechanisms, pancreatic

cancer, radiofrequency ablation, AM, lattice structures, and design optimization.

1.1.1 Compliant Mechanisms

A compliant mechanism (CM) gains its motion through the deformation of a material as

opposed to traditional hinges or joints [1]. Because the mechanisms do not use traditional hinges

for movement, they have several natural benefits such as reduced part count or assembly costs,

energy absorption in deflected members, and reduced wear due to friction. In addition, there are

some unique advantages for surgical CMs. In surgical applications, compliant mechanisms do not

require lubrication, lack wear, reduce the number of parts, and can improve mobility or dexterity

[2,3]. A number of surgical tools have been designed which exploit the unique advantages of

compliant mechanisms [4–7]. An example of a compliant surgical forceps is shown in Figure 1.1.

Figure 1.1 Compliant surgical forceps reduce the need for complex hinges on a small scale, lack

debris from wear, and do not require lubrication, adapted from [8]

Page 32: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

3

One major challenge associated with CMs is that they often consist of complex structures

making them difficult to manufacture. For this reason, CMs were often planar mechanisms to

make them manufacturable. Recently, with the advancements of AM, there are new opportunities

to manufacture complex 3D CMs. Because AM also enables part consolidation and the ability to

manufacture more complex parts than traditional subtractive processes, it becomes a natural fit to

consider AM for building CMs.

With the ease of accessibility to polymer AM machines, polymer CMs have become very

common [9–12] , especially in the area of multi-material polymer AM [13–15]. Metal AM CMs

are much less common than polymer CMs for multiple reasons. One major challenge is that the

cost of metal AM remains high, making it less accessible than polymer AM processes. A second

reason is because much of the focus of AM is on structures designed to be as stiff and lightweight

as possible. For metal AM CMs, Merriam et al. [16–19] have developed titanium compliant

mechanisms for space applications. Other examples of AM CMs found in literature include

functionally graded superelastic compliant joints [20] and a high precision compliant parallel

mechanism [21]. The potential for CMs or other metamaterials designed specifically for metal

AM is largely unexplored with limited guidance or feedback from the manufacturing process to

direct the design. Two examples of compliant mechanisms designed for AM are shown in Figure

1.2. With the increased manufacturability of CMs, there are opportunities to improve

performance of components in a variety of different industries. One area where CMs excel is in

the medical industry because they do not require lubrication or wearing surfaces at the joints. In

this dissertation, the motivating application of CMs is a surgical tool required for treating cancers,

such as pancreatic cancer.

Page 33: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

4

Figure 1.2 Examples of multi-material (left) and metal (right) compliant mechanisms designed for

AM, adapted from [13,19]

1.1.2 Pancreatic Cancer

Pancreatic cancer is the fourth most deadly cancer for men and women in the United

States [22]. A major cause of the poor prognosis is the difficulty of detection resulting in nearly

85-90% of patients being diagnosed at advanced stages of the disease [23]. Pancreatic cancer is

diagnosed into one of four stages based on the location of the tumor and how much it has spread.

Often the first symptoms a patient feels are pain and discomfort due to an enlarged tumor.

Depending on the location of the tumor, additional symptoms may include weight loss or

jaundice. At this point, when the symptoms become noticeable, the tumor has likely impacted

surrounding organs and vessels and possibly spread to other areas of the body. While universal

screening has been suggested to increase earlier detection, it has not yet been proven effective

[24]. In addition, pancreatic cancer is more common in older patients with poorer health than in

young healthy patients.

Due to the late diagnosis and poor health of many pancreatic cancer patients, the average

five year survival rate from the time of diagnosis is only 7% [22]. A comparison of ten of the

most common cancers and their five-year survival rates are shown in Figure 1.3. A patient

Page 34: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

5

diagnosed in late stages of pancreatic cancer is often limited to palliative care to reduce pain and

discomfort because of the spread of the disease. The poor health of many patients also limits the

treatment options available. While earlier detection of pancreatic cancer remains a challenge,

incidental detection is increasing due to improved patient imaging. A study by Lee et al. [25]

found incidental pancreatic cysts in 13.5% of the adult population using retrospectively viewed

MR images. Another similar study determined that 60% of the pancreatic cysts in patients older

than 70 are likely to be malignant [26]. With the increase of incidental pancreatic cysts being

discovered at earlier stages, more patients will be seeking curative treatment.

Figure 1.3 Five year survival rates for the 10 most common cancers [22]. Pancreatic cancer is the

lowest with an average fiver year survival rate of 7%.

When treating cancer, interdisciplinary teams of doctors meet to review a patient’s case

and determine which forms of treatment should be used to best combat their disease. A treatment

plan is generated based on various information available to the doctors including the cancer type,

† Includes renal pelvis

‡ Includes intrahepatic bile duct

Page 35: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

6

location, and stage, as well as the health of the patient. Medical imaging plays a critical role in

helping doctors plan an appropriate treatment by visualizing potential risks associated with the

disease and side-effects from any action they take. There are a number of treatment modes for

cancer, the most common being surgical resection, chemotherapy, radiation therapy. In addition,

Knavel et al. [27] provides a review of various other ablative technologies, such as cryoablation,

radiofrequency ablation, microwave ablation, and high intensity focused ultrasound. These teams

of doctors seek to combine the benefits of multiple forms of treatment while minimizing the

negative effects on a case by case basis.

The most effective form of treatment for non-metastatic pancreatic cancer is surgical

resection followed by an adjuvant or secondary treatment. During surgery, a doctor removes the

tumor and layers of surrounding tissue until a negative margin can be developed around the

tumor. When the full tumor cannot be removed due to the size or location, the surgeon seeks to

remove large portions to reduce the overall size. Surgery allows the doctor to visualize the tumor

and surrounding tissues as well as inspect other areas of the body where cancer may have spread.

A secondary treatment such as chemotherapy or radiation therapy is typically used to destroy any

cancer cells that may have been missed or further reduce unresected portions of the tumor. Chen

and Prinz [28] and Werner et al. [29] reviewed many studies which have discussed the effects of

various treatments on long-term survival of pancreatic cancer patients. While multiple studies

reviewed by Chen and Prinz [28] showed that resection followed by chemotherapy tends to

increase the survival of patients, the patient population for each study was often very small

resulting in unconvincing results of the benefit of chemotherapy following resection. One of the

strongest studies reviewed by Chen and Prinz was published in 2004 by Neoptolemos et al. [30].

This study included 289 patients and presented convincing results that surgical resection of

pancreatic cancer followed by chemotherapy increased survival. Werner highlights a number of

Page 36: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

7

recent studies which show improved long term survival for patients who received an adjuvant

therapy following surgery [31–34].

While the combination of surgery with chemotherapy is the most effective form of

treatment, many patients with non-metastatic pancreatic cancer are not healthy enough to undergo

this complex surgery and recovery. Minimally invasive approaches, such as laparoscopy or

endoscopy, are available to treat many cancer patients who are ineligible for open surgery.

Laparoscopy can be used for delivery of chemotherapy, radiation therapy, and various forms of

ablation technologies. One common minimally invasive approach which has shown promise is

radiofrequency ablation (RFA).

1.1.3 Radiofrequency Ablation

Radiofrequency ablation (RFA) is a common mode of cancer treatment that has been

used since the early 1990’s. The effect of heating due to radiofrequency waves was first reported

in 1891 by D’Arsonval [35], followed by the explanation of the physical principles in 1976 by

Organ [36,37]. The radiofrequency waves being passed through tissue were shown to produce

oscillation in the ions causing frictional heating in the tissue. Later in the 1990’s McGahan et al.

[38] and Rossi et al. [39] first attempted RFA in the liver. RFA is now a common treatment

method that has been applied in various locations throughout the body including the lungs, liver,

kidneys, and pancreas [27,40].

RFA is performed by inserting one or more electrodes into a tumor and a large counter

electrode is placed on the exterior of the patient. A radiofrequency current applied to the electrode

causes frictional heating of the tissue immediately surrounding the electrode where the

concentration of the electric field is highest [41]. The frictional heating is a result of the

alternating current causing oscillation of the ions near the electrode. The primary mode of heat

Page 37: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

8

transfer is conduction through the tissue, raising its temperature and destroying the cells. The cell

destruction is time and temperature dependent. The target tissue temperature for RFA is between

50-100°C [41]. For temperatures below 50°C, the cells require a long exposure time to be

destroyed and at temperatures greater than 100°C, the cells char, increasing their resistance and

limiting the effectiveness of the current flow. In the range of 50-60°C, reported values for the

exposure time required to cause cell death vary greatly in literature [27,40–43]. However, there is

agreement in that for cells heated to temperatures between 60-100°C, cell death is nearly

instantaneous [40,42,44–46]. The region of cell death surrounding the electrode is referred to as

the ablation zone. The study of the effects of temperature on cells is called thermal therapy. Many

studies have been reported in literature about the effects of hyperthermia on cells [42,43].

Dewhirst et al. [47] provides an summary of previous work and the focus of thermal therapy

moving forward.

RFA is primarily limited by three factors: tumor location, tumor geometry, and tracking

the treatment zone. First, location-based limitations occur when the tumor is near other vital

structures or organs, blood vessels, or in a region of the body which is difficult to access using

traditional delivery methods. Effective RFA treatment requires the entire tumor plus a small

margin around the periphery to be fully heated by the electrode. Because RFA relies on thermal

therapy to sufficiently heat and destroy tumor, if there are vital structures or organs adjacent to

the tumor then they may be negatively affected by the heating. In particular, it is difficult to

develop a negative margin around the tumor to ensure all cancerous tissue is destroyed without

damaging healthy surrounding tissue. If the tumor is located near a large blood vessel (diameter

larger than 3mm), the blood vessel counteracts the effects of RFA by acting as a significant heat

sink [48,49]. As the heating from RFA is primarily based on conduction, the blood vessel quickly

draws away the heat, preventing the tumor from heating sufficiently to receive permanent

damage. A final challenge with the tumor location is simply accessibility for surgical tools. In

Page 38: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

9

certain regions such as the pancreas or other areas of the upper abdomen, it is difficult to find an

access point for the RFA electrode.

The second limitation is due to the geometry of the electrode and tumor. These

limitations occur when the heating profile generated by the electrode does not match the tumor

geometry. For example, if the heating profile is cylindrical and the tumor is spherical, there is a

mismatch which either damages healthy tissue or does not destroy all the cancer; see Figure 1.4.

To overcome this limitation, a few different electrode configurations are commercially available

and have been reported in literature [41,50,51]. Multiple reviews of single electrodes and multi-

electrode systems are available such as those by Hong and Georgiades [41] and Mulier et al. [51].

Hong and Georgiades [41] gave a review of current single electrodes and electrode systems. The

most basic shape of RFA electrodes is cylindrical which typically generates an ellipsoidal

ablation zone. The exposed length of the electrode can be adjusted or multiple cylindrical

electrodes can be inserted in an attempt to alter the geometry of the ablation zone. Hong and

Georgiades [41] also describe complex expanding electrodes that broaden the ablation zone

through multiple tines that deploy or expand into tissue as the electrode is inserted. Mulier et al.

[51] includes a review of current electrodes and multiple electrode systems as part of his proposal

for standardized terminology of RFA devices; see Figure 1.5 [51]. In his review of RFA for

pancreatic adenocarcinoma, D’Onofrio et al. [52] describe how straight electrodes are used for

cylindrical-like tumors and deployable electrodes are used for spherical-like tumors. Even with

several configurations of electrodes available to try to match the ablation zone and tumor

geometry, it is hard to ensure a negative margin is developed around the tumor because the

ablation zone is difficult to predict and the tumor geometry is unique to each patient.

Page 39: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

10

Figure 1.4 (Left) Diagram of radiofrequency ablation showing the mismatch that can occur between

a tumor and ablation zone and (Right) thermal-imaging of radiofrequency ablation showing the

ellipsoidal temperature profile around a straight electrode, adapted from [41]

Figure 1.5 Basic electrode types and several commercial models of radiofrequency ablation

electrodes described by Mulier et al., images adapted from [51]

Tumor

Ablation

Zone

Page 40: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

11

The third limitation is the ability to track and get feedback on the treatment zone shape

and size. Three primary methods of tracking of the ablation zone during radiofrequency ablation

have been proposed: ultrasound (US), computed tomography (CT), magnetic resonance imaging

(MRI). US provides real-time tracking of the ablation zone but can be limited due to gas bubble

artifacts which blur the image during ablation and it is often associated with underestimation of

the actual ablation zone [53–55]. CT imaging has been correlated with improved ablation zone

prediction but is less available than US and often is unable to provide real-time feedback [56].

MR imaging is expensive, time consuming, and cannot be performed real-time [57]. In current

practice, radiofrequency generators often monitor the impedance or temperature at the electrode

to provide automated control power delivery and ablation zone generation. Other research groups

are currently pursuing impedance- or temperature-based automated control to predict ablation

zones [58]. Because it is difficult to predict and monitor the treatment zone, recurrence or

multiple treatments are common following the initial RFA.

The result of the current limitations of RFA is incomplete tumor ablation. One reason

surgical resection is so effective is because the doctor is able to visibly remove the tumor

followed by a few layers of extra tissue around the periphery to ensure that no tumor cells have

been missed. If no cancer cells can be found at the outer edge of the tissue where the tumor was

located, the patient has a negative margin resection. With RFA it is very difficult to ensure a

negative margin is developed around the tumor because of the limitations described above. Using

a specially shaped electrode to create an ablation zone which matches the tumor geometry and

combining RFA with other treatment modalities becomes increasingly important in developing

the desired negative margin around the tumor and preventing reoccurrence.

To understand RFA in tissue, numerous finite element models have been created. RFA

models have been applied in thyroid tissue based on MRI reconstructed models and also in liver-

tissue using a hyperbolic bioheat equation and a new-voltage calibration method [59,60]. Various

Page 41: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

12

ablation models have been used to study the effects of tissue properties such as electrical

conductivity and blood perfusion rate [61–66] or the heat sink effect from blood vessels [67–69] .

Duan et al. [64] and Huang and Chui [66] both present unique work for treatment planning that

statistically predicts the ablation zone by varying tissue properties [70]. Tungjitkusolmun et al.

[67] presented a 2D and 3D model of ablation using the Rita Medical Systems Electrode and

explored the effects of ablation near blood vessels. Other models focus on specific commercially

available electrodes [44,71–74] or automatic power delivery methods [58,60,75]. Chang [61]

focused on the effects of the blood perfusion rate and temperature dependent electrical

conductivity. Later, Chang and Nguyen [71] studied how isotherms can be used to predict the

ablation zone. In addition, Altrogge et al. [73] studied the optimization of the placement of a

commercial electrode to increase the efficiency of RFA. Another model by Lim et al. [68]

focused on improving efficiency by considering alternative input waveform patterns. While many

models have been developed to understand RFA, the use of these models within a formal design

optimization of probe tip geometry to create a specific ablation zone has not been considered.

Up to this point, models available in literature study commercially available electrodes

and have not explored the use of such models in conjunction with a systematic design

optimization tool. For a given tumor geometry, these models could be used with design tools to

direct the shape and size of an electrode to match the ablation zone and tumor geometries.

1.1.3.1 Endoscopic RFA

Due to the challenges with placing an RFA electrode percutaneously in certain regions of

the body and with the intent of making surgery even less invasive than open surgery, recently

endoscopic RFA has been explored. For example, in the case of pancreatic tumors, the tumor is

Page 42: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

13

often inaccessible because of the surrounding organs and location of the tumor in the pancreas.

Endoscopic RFA offers a unique approach that allows access for insertion of an electrode.

Endoscopic ultrasound-guided radiofrequency ablation (EUS-RFA) is a relatively new

approach which allows for increased access to certain areas in the upper abdomen such as the

pancreas. It has been tested in several animal models and clinical trials. Varadarajulu et al. [76]

and Gaidhane et al. [77] performed EUS-RFA in pigs on the liver and pancreas respectively.

Varadarajulu found that after seven minutes the ablation was obtained and the electrode could

easily be retracted. The pigs remained stable and without significant damage to surrounding

organs after producing a 2.6cm diameter ablation zone. Gaidhane found similar promising results

in 5 Yucatan pigs causing minimal pancreatitis and concluded that EUS-RFA is a safe procedure.

Following their successful results, Pai et al. [78] conducted clinical trials on eight human patients

and found the EUS-RFA treatment safe, with minimal risks to patients. They found a range of

results from complete resolution to 50% reduction in diameter. Other recent studies and reviews

have continued to show promising results [79–84]. While the trials conducted have been effective

and safe, repositioning the electrode and repeated applications of RFA in a single session were

required to generate a large ablation zone and match the geometry of the tumor [78].

While EUS-RFA has been shown to be safe and increases accessibility to other areas of

the body, there are only two EUS-RFA electrodes which have been tested in clinical trials, the

Habib EUS-RFA catheter (Emcision Ltd, London) [77,78,85] and the cryotherm probe (ERBE

Elektromedizin GmbH, Tübingen, Germany) [86,87]. Each electrode is a straight cylinder, which

generates an elliptical ablation zone similar to that shown in Figure 1.6. There is a need for

specialized electrode designs that can reduce the geometrical mismatch that often occurs in

treatment of pancreatic cancer with RFA.

The first systematic optimization method described in the dissertation is focused on a

shape matching method RFA to design specialized electrodes. The optimization method is

Page 43: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

14

applied to a novel EUS-RFA electrode design that is proposed in this work. The second

systematic optimization is for tailored mesostructures or lattices. This optimization includes AM

considerations and process constraints.

1.1.4 Additive Manufacturing

AM has seen significant growth in recent years due to the increased interest from a

variety of industries such as automotive, aerospace, medicine, biological systems, and food

supply chains due to several of its unique advantages [88–90]. Traditional subtractive

manufacturing processes begin with a solid piece of material and progressively remove or

subtract material to achieve the final form. AM refers to seven different manufacturing processes

that involve adding material layer-by-layer: material extrusion, powder bed fusion, vat

polymerization, material jetting, binder jetting, sheet lamination, and directed energy deposition.

Abdulhameed et al. [91] reviewed all seven processes in depth to identify advantages and

challenge with each process. The layer-by-layer approach of AM enables generation of parts that

were previously impossible or too complex to manufacture and reduces material waste compared

to subtractive manufacturing. In reality, with AM, the cost or manufacturability of a part is not

directly related to its complexity as with traditional subtractive processes, where the number of

steps to create the part or number of different tools drive the cost of the part. While AM is not

without its limitations, there is arguably an inherent “free complexity” in design based on the fact

that a part is built layer by layer with material deposited where it is needed. The concept of free

complexity combined with the increased capability to manufacture complex internal geometries

has led to a variety of research studies into AM for applications such as compliant mechanisms,

discussed in Section 1.1.1.

Page 44: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

15

In their review of AM, Gao et al. [88] describes how AM, also known as desktop

manufacturing, rapid manufacturing, rapid prototyping, was most commonly used for quick

demonstration pieces but has progressed toward end use products. Gao et al. [88] outlines several

pros and cons associated with AM that are often lumped into two groups called opportunistic and

restrictive. From the opportunistic side, Gao highlights increased design flexibility, free

complexity, part consolidation and/or printed assemblies, and cost efficiency for low production

quantities [88]. Many examples of opportunistic AM exist in literature including non-assembly

builds [10,92], customized parts [93–95], part consolidation [96,97], and multi-functional parts

[98]. A few examples of multi-functional parts are embedded components or actuators [99–101]

and multi-material printing [102,103].

The restrictions or challenges with AM include high cost for mass production, limited

scalability due to increased build time or loss of resolution, part consistency and reliability, or

AM standardization and intellectual property [88]. Two of the biggest limitations to AM are the

cost for mass production and part reliability or consistency. With other processes such as

injection molding, after molds are produced, the cost per part decreases as the production quantity

increases. In contrast, for AM there is a much small upfront cost for machines but the cost per

part does not decrease with production quantity. As the technology is still in its infancy for end

use products, there are still many concerns over the reliability and consistency of parts. In

particular for custom parts, an important question is, how can the safety and quality of every

product be validated if each product is customized? To improve part reliability a major area of

research is in-situ monitoring methods to improve consistency and detect flaws [104–111]. While

many industries rely on driving down cost by mass production through consistent and reliable

processes, AM is continually improving to reduce build time and improve consistency and

dependability.

Page 45: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

16

1.1.4.1 Metal Powder Bed Fusion

Several divisions in AM exist based on material, process, and energy type. For metals,

there are two major processes: powder bed fusion (PBF) and directed energy deposition (DED).

For DED there are two material feedstock methods: wire and powder. In DED, the energy source

heats the part generating a melting pool and a wire or powder is fed into melt pool. Either the part

itself or the energy source and feed material are moved generating layers of added material. These

deposited layers progressively build the part. PBF is performed by spreading a thin layer of metal

powder on a metal build plate, then using a laser or electron beam to melt the powder to the build

plate. The melted powder forms to the first layer of the final part. The build plate is then lowered

and a new layer of metal powder is spread across the previous layer. Again, the powder is melted

to the previous layer and the process is repeated, slowly building the metal part layer by layer.

Schematics of PBF, powder feed DED, and wire feed DED systems are shown in Figure 1.6.

In the review of metal AM, Frazier [112] reports that several challenges exist for metal

AM, such as the need for monitoring to improve part consistency and increased material selection

through alloy development. One challenge with consistency is the numerous thermal cycles that a

part experiences causing variations in microstructure and deformation of the part. Even with the

need to improve consistency and develop materials specific for AM, Frazier [112] reports that the

static mechanical properties are similar to those of conventionally processed parts. Specific to

PBF, the speed, accuracy, and cost need to improve and the material processing-structure-

property relationships need to be understood to improve the technology [113]. By reviewing

published works reporting microstructure and material properties, Herzog et al. [114] provides a

thorough review of metal AM processes, linking processing parameters and material properties in

various types of metal.

Page 46: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

17

Figure 1.6 Metal AM system types include powder bed fusion and wire/powder feed directed energy

deposition, typically using either a laser or electron beam energy source, adapted from [112]

1.1.4.2 Design for Additive Manufacturing

Design for AM (DfAM) includes the theories and principles behind preparing parts to be

manufactured using AM, taking full advantage of the unique opportunities and avoiding the

restrictions. A common misconception exists in that there is complete design freedom with AM,

but the design freedom of AM is in fact not limitless. In 2016, the U.S. Department of Defense

established a roadmap for AM including four focus areas: Design, Material, Process, and Value

Chain. In the design focus area, specifically DfAM tools that “streamline [the] design process,

Powder Bed System Wire Feed System

Powder Feed System

Page 47: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

18

reduced cycle time, and [create] higher performance products” [115] were highlighted as future

tasks. Because the manufacturing process is very distinct, the design rules for traditional

manufacturing processes are naturally different than those for AM. In addition, because the

design freedom is increased, DfAM tools are needed to help guide a designer in enhancing

product performance over traditional designs. With DfAM tools, parts can be tailored for design

requirements ranging from biologically inspired effective modulus properties to improve

orthopedic implants to multifunctional structures [116–118]. While there is a wide range of

possible applications, DfAM optimization tools to design these structures may be one in the same.

DfAM is generally separated into two categories: opportunistic and restrictive DfAM.

Describing opportunistic DfAM, Gibson et al. [119] organized it into four groups: shape

complexity, hierarchical complexity, material complexity, and functional complexity. These

groups were defined as follows:

• Shape complexity – ability to create almost any freeform shape,

• Hierarchical complexity – ability to design across multiple length scales ranging from

the grain structure and orientation, or micro-level, internal geometric complexity, or

meso-level, and the general object shape or form, or macro-level,

• Material complexity – ability to create parts one point at a time, mixing and blending

multiple materials,

• Functional complexity – ability to access the entire 3D geometry during the

manufacturing process enables integration of multiple functional parts through

printed assemblies or embedded components.

Using a slightly different method of categorization, a review by Thompson et al. [120]

provides a discussion on numerous areas of both opportunistic and restrictive DfAM; see Table

1.1 for a complete list of the topics addressed. Thompson et al. [120] covers everything from the

ability to print full color to designing the microstructure to the regulatory limitations of the

relatively new technology. Gibson and Thompson both acknowledge that many of these DfAM

topics are in their infancy and merit further study to maximize their capability of minimize their

Page 48: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

19

limitation [119,120]. For example, on the opportunistic side, meso-level DfAM enables further

enhancement to maximize performance by reducing weight or tailoring properties, but no fixed

guidelines or design rules have been established. Much of the meso-level design remains based

on designer intuition as opposed to concrete design rules or methods. On the restrictive side,

minimum feature size and the need for support material are established as necessary

considerations in DfAM, but there is ongoing research in ways to reduce the negative impact of

these topics. In an attempt to incorporate a variety of DfAM topics, there are several DfAM

frameworks that have been proposed.

Table 1.1 A wide range of opportunistic and restrictive DfAM topics exist for which there is ongoing

research

Opportunistic DfAM

Macro-Level Meso-Level Product-Level

• Material Choice

• Full Color

• Freeform geometry for

art/aesthetics

• Internal freeform geometry

• Topology optimized macro-

structure

• Custom fit or mass

customization

• Custom material, metallurgy,

microstructure

• Custom surfaces, textures,

porosity

• Lattices, trusses, cellular

materials

• Multi-material

• Functionally graded

• Metamaterials

• Part Consolidation

• Embedded objects

• Printed Assemblies

Restrictive DfAM

CAD and Digitization Discretization, Directionality,

Build Orientation, and Supports

Process Characteristics and

Machine Capabilities

• Complete model

• Inadequate for organic shapes

• Limited meso-structure

design

• Surface roughness dependent

on orientation

• Supports for overhangs

• Boundaries between layers

• Minimum Feature Size

• Accuracy

• Build Volume

• Compatible materials

• Process specific design rules

Material properties and

processing

Metrology and quality

control

Maintenance, repair,

recycling

External and regulatory

constraints

• Need materials for

AM

• Properties change

with processing

• Voids, porosity,

grain characteristics

• Internal geometries

• Disassembly of

printed assemblies

for maintenance

• Relatively new

technology

• Inherent variability

Page 49: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

20

DfAM frameworks are a set of guidelines or methods to aid designers in creating efficient

and successful parts for AM that take advantage of the opportunities available while accounting

for the limitations. For example, ASTM 52910 is a standard which very broadly addresses

guidelines for DfAM [121]. In literature there are many research groups trying to provide the

necessary design guidelines for DfAM [122]. After a review of 27 publications on DfAM for

decision making, Laverne et al. [123] describe that the articles could be categorized by their focus

on opportunistic DfAM, restrictive DfAM, or dual DfAM, a combination of both. To take full

advantage of AM and have a successful build both opportunistic and restrictive DfAM must be

considered, yet Laverne et al. [123] found that only 36% were dual DfAM. Yang and Zhao [124]

provide a review of AM design theory, describing several general design considerations for AM,

DfAM frameworks, systematic design tools for AM, and makes recommendations for future

research areas in DfAM. In reviewing DfAM frameworks, they were grouped into two categories:

AM-enabled structural optimization and DfAM methodologies.

AM-enabled structural optimization covers the novel shapes and structures that can be

created at both the macro and meso-levels. This includes various forms of topology optimization

to generate optimized structures that would otherwise be difficult to manufacture. Additional

detail on topology optimization is covered in Section 1.1.6.2. Of these DfAM optimization

methods, Yang and Zhao [124] describe the benefits of combining functional requirements

through objective functions to optimize functional surfaces or complex designs. While there are

clear advantages to such approaches for generating an optimal design, there are also limitations

with the ability to include other unique opportunistic features of AM. Later in their discussion of

systematic design tools for DfAM, Yang and Zhao [124] describe design tools that allow a

designer to alter the interior volume or meso-level design. While there are several commercial

tools for optimizing the interior volume such as OptiStruct [125] and Netfabb Ultimate [126],

they perform with relatively limited options, not providing full control for a designer to specify

Page 50: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

21

various objective functions, but primarily focus on stiff and light weight solutions. In their review

of AM design theories and methodologies, Yang and Zhao [124] found that there is a need for a

design framework that “initializes design from the perspective of functionality achievement” and

a “method to better synthesize functional requirements and process knowledge simultaneously.”

Regarding DfAM, Doubrovski et al. [127] studied the relationships between structure and

performance as well as optimization approaches. Nearly half the articles reviewed by Doubrovski

et al. [127] related the structure to strength- and stiffness-based performance. Aside from

strength and stiffness, the other articles were spread throughout other performance objectives

including thermal, compliance, dynamic, and visual. With early interest in AM from the

automotive and aerospace industries, much of the focus of designing AM parts has been

concentrated around weight reduction and creating stiff light-weight structures. Little work was

reported in the area of tailored compliance or thermal properties at the mesostructural level.

Doubrovski et al. [127] conclude that “One of the promising possibilities of AM is the complexity

of geometry… combined with methods for optimization” and to take full advantage of AM in the

product development process systematic design tools have to be developed [127]. While it is

challenging to create a general set of DfAM guidelines or design tools that apply across all AM

processes, there are many researchers exploring design rules specifically for metal AM.

In 2017, a report from the Fraunhofer institute detailed the design of 7 different

components for AM. One of the purposes of detailing these 7 different industry-relevant case

studies was to highlight DfAM guidelines and limitations with current metal AM technology.

Over 20 different metal AM-relevant industry standards were identified and considered during the

different case studies. Only three of these standards deal directly with DfAM (ISO/ASTM 52910,

VDI 3405 Part 3, VDI 3405 Part 3.5). Few other general metal AM guidelines or standards exist

for metal AM though it is a rapidly evolving field [128–132].

Page 51: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

22

A few of the primary manufacturing limitations and considerations for metal PBF are the

required support structures, overhangs, and powder removal. Metal support structures are

required in PBF for three purposes: anchoring the part to the build platform, conducting heat

away from the part, and supporting overhanging features. In metal AM, the part undergoes large

changes in temperature caused by reheating for several subsequent layers [114]. During

processing, the resulting thermal stresses cause parts to warp upward away from the build plate.

The warping occurs when hot upper layers cool and contract while lower layers do not contract,

causing the upward warping. Figure 1.7 shows a basic schematic of the upward warping. The

support structures in metal AM help to anchor parts to the build plate and also prevent warping.

Heat buildup causes increased surface roughness and residual stresses in the part. Because solid

metal structures have higher conductivity than metal powder, support structures conduct heat

away from the part better than powder. Similar to other AM processes, overhanging features

require support material. As a rule of thumb, overhangs of less than 45° require support structures

to both support the melted material and anchor it to the build plate. In recent years, the minimum

build angle has continued to decrease, including some companies, such as VELO3D, that boast

support-free designs. Specific to PBF processes for AM is the requirement that the geometry of

the design allow for powder removal, which limits enclosed cavities and other small features that

may trap powder. It occurs due to geometrically tight channels that are difficult for powder to

flow from as well as due to dross or excess powder being picked up because of poor thermal

management of the process. Trapped powder increases the weight, cost, and safety of a

component. Rather than empty cavities in a part, cavities filled with powder adds cost through

unrecoverable powder and weight without any structural integrity. Additional safety concerns

come into play with powder that could eventually find its way out of the part posing health risks

if aspirated. Processes to remove powder from a component include vacuuming, compressed air,

Page 52: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

23

tapping the part, ultrasonic vibration, and mounting it on a multi-axis turntable for rotating the

part and allowing powder to flow out in various directions.

Figure 1.7 Example of warping due to thermal contraction

1.1.5 Lattice Structures

One major area of research for AM, with significant potential for enhanced performance,

is the ability to design at the mesoscale [133,134]. Mesoscale design, also known as metamaterial

design or design of cellular structures, is designing the functions or properties of a material as a

direct result of its mesostructure geometry as opposed to the raw material. The mesoscale is an

intermediate level between macroscale and microscale, including features at length scales of

approximately 0.1–10mm [135]. For example, in a beehive, the macroscale of a component is the

overall structure or topology of the outside, the mesoscale is the honeycomb structure within the

outer shell, and the microstructure is the makeup or material state of the individual walls of the

honeycomb. The ability to design and manufacture the macrostructure or overall topology of

complex parts is a natural benefit of the “free complexity” of AM. As the accuracy and precision

of AM processes improve, increased complexity at the mesoscale and even microscale may be

designed and manufactured without direct increase in cost of the part. Common examples of

mesostructural design with traditional manufacturing processes are honeycomb structures and

metal foams; see Figure 1.8. Honeycomb and cellular structures have been observed in nature for

Page 53: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

24

thousands of years but it wasn’t until the early 1900’s that they were first used in structural

applications such as airplane wings and buildings. Since then, many researchers have studied and

tried to replicate cellular structures, but for many years the complexity of these microstructures

was limited [136–139]. For organized structures, such as the honeycomb, only limited designs

could be manufactured efficiently, if at all. Several examples of traditionally manufactured lattice

structures have been researched and developed by Haydn Wadley [140–157]. More variety can be

found with metal foams than organized structures but they suffer in applications where a

repeatable structure is needed, due to the stochastic nature of their fabrication. For many years,

researchers have been creating and analyzing unique cellular structures analytically or with the

finite element method. Now, there is a renewed interest in mesoscale design due to advances in

AM technology that allow for these complex structures to be manufactured in 3D.

Page 54: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

25

Figure 1.8 Various forms of cellular solids including a (top) 2D honeycomb, (middle) 3D open-cell

foam, and (bottom) 3D closed-cell foam, adapted from [136]

Though a variety of names have been used in the past for cellular structures, lattice

structure is the terminology used throughout this dissertation. Lattice structure refers specifically

to an organized repeating mesostructure for which the most basic repeating unit is called a unit

cell [117,158,159]. Lattice structures offer improved strength to weight performance over solid

components by maintaining structural integrity while decreasing weight [136]. Helou and Kara

[160] give a comprehensive review of lattice structure literature and specify that AM processes

have an “immense impact on the form and functionality of the specimen being produced.”

Though there are other methods to create lattice structures, the focus of this dissertation and

literature review is on lattice structures fabricated with AM. Lattice structures have been used for

Page 55: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

26

a variety of applications including light-weighting [161], functionally graded structures [162],

tissue engineering scaffolds [163–169], medical implants [170–181], actuation[182], heat transfer

[118,183–185], reinforcement [186], and energy absorption [187–193]. Figure 1.9 shows an

example of various medical implants, each with a different approach to lattice structure design.

Because lattice structures may be incorporated into such a wide variety of applications, there is a

need for further research exploring how to design or select an appropriate unit cell topology for

the application.

Page 56: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

27

Figure 1.9 Lattice structures have been proposed for a variety of medical implants with the intent of

improving performance and functionality. Due to the variety of lattice structures that can be used in

such applications, there is a need for further research to understand how to select or design an

appropriate unit cell for the intended application. Images adapted from A) [179], B) [177], C) [181],

D) & E) [194].

Designing a lattice structure within a component is typically performed by separating the

process into two distinct steps: unit cell generation and lattice population. Unit cell generation is

commonly based on three different approaches as described by Tao and Leu [195], (1) basic

primitive shapes, (2) mathematical equation based surfaces, or (3) topology optimization. Basic

primitive unit cells are generated by combining common shapes and different Boolean operations,

A) B)

C)

E)

D)

Page 57: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

28

often in ways that mimic common crystal structures seen in nature [196,197]; see Figure 1.10. A

few examples of common basic primitives are the body-centered cubic (BCC), face-centered

cubic (FCC), octet-truss. For modeling unit cells with mathematical equations, three-dimensional

surfaces are generated using special formulations. Examples of unit cells from mathematical

equation-based surfaces are the gyroid or another example shown by Tao and Leu [195], whose

shape and unit cell density may be adjusted by two parameters; see Figure 1.11. With infinite

possibilities for the topology of the unit cell, several naming conventions or classifications have

been proposed [198–200]. Unit cell density refers specifically to the volume fraction of the filled

space within the bounding volume of the unit cell. For example, 50% dense would be a unit cell

sized such that half of the volume within the bounding box is full and 100% dense would mean

the bounding box is filled completely. Topology optimization is a developing research area for

unit cell generation and is covered in Section 1.1.6.2.

Figure 1.10 Unit Cells are often generated based on Boolean operations with primitive shapes such as

these examples, adapted from [195]

Page 58: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

29

Figure 1.11 Unit cells may be generated using implicit mathematical equations to represent a surface.

In this particular case, the surface is defined at F(x,y,z) = 0 with parameters "a" and "b" used to

tailor the shape of the surface; adapted from [195].

To understand the physical behavior of different unit cells and their properties, numerous

researchers are developing analytical [178,201–206] and finite element methods [202,204–210] to

predict the properties of different unit cell types and sizing, and performing experimental

validation. In current literature for metal AM, unit cell properties may be found for cube

[178,203,211], rhombic dodecahedron [178,211,212], rhombicuboctahedron [178,204,211], BCC

[178,208,213–216], FCC[208,213], diamond[211,217–219], gyroid [213,220–223], pyramidal

[190], octahedral [178], and several other variations on these. A complete set of sources is

described as part of Chapter 6. Dong et al. [116] performed a review of modeling approaches,

comparing the advantages and disadvantages of each approach. Among their primary findings

was that lattice structures are promising with their potential to be multi-functional materials in

various engineering applications.

𝐹(𝑥, 𝑦, 𝑧) = cos(2𝜋𝑥) + cos(2𝜋𝑦) + cos(2𝜋𝑧)+ 𝑎(cos(2𝜋𝑥) cos(2𝜋𝑦) + cos(2𝜋𝑦) cos(2𝜋𝑧)+ cos(2𝜋𝑧) cos(2𝜋x)) + 𝑏

Page 59: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

30

Lattice population or patterning throughout a part is accomplished using either Boolean

operations or conformal methods. Basic Boolean operations are the most common and simplest

method for populating a unit cell within a part. The unit cell is repeated into a lattice block and

the edges of a part or volume are used to trim the lattice. Often only partial unit cells remain near

the edges of the part resulting potential weak points. Conformal lattice population methods

consist of modifying the lattice block to match the surface geometry of your part. A comparison

of conformal and non-conformal lattices is shown in Figure 1.12. Rosen and his students have

performed extensive work into conformal lattice methods and design processes for generating

lattice structures [159,224–227].

Figure 1.12 Two common methods for lattice population are (top) uniform in which the basic lattice

is not modified based on the component shape and (bottom) conformal in which each unit cell may be

deformed to follow the surface of the component, adapted from [224]

After these two steps of generating a unit cell and populating the lattice structure, often

the lattice is altered by a sizing optimization, changing thickness or strut diameters to provide

higher stiffness in certain areas than others. Several other approaches have been proposed in

Uniform lattice

Conformal lattice

Page 60: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

31

recent years which blend together these steps: unit cell generation, population, and sizing, often

all mixed with an optimization routine [228]. For example, Brackett et al. [229] and Arabnejad et

al. [230] use a density mapping to size and position their lattice structure. Brackett et al. [229]

uses a gray-scale image of the component stress to generate the density mapping for the lattice

structure. The gray-scale stress image is generated by finite element analysis and then dithering

the image; see Figure 1.13. A lattice structure is populated through the part with small stiffer

trusses in regions of high stress or a high density of points. Arabnejad et al. [179,230–232]

generates a lattice structure inside a hip implant using the density mapping from a CT-scan. The

densities from the scan are correlated to a particular unit cell density to fill the part; see Figure

1.14. Graf et al. [233] and Chang and Rosen [234] use similar approaches for generating and

sizing a lattice. First, they begin with a basic lattice and either add, remove, or resize struts on

each unit cell based on the type of loading and amount of stress in that region of the component.

The stress is found using finite element analysis. High stress regions of the component under load

are correlated with increasing unit cell density; see Figure 1.15. While these methods have proven

useful, they lack the full customization and ability to optimize performance that is possible for

AM. While they may serve to lightweight a component and improve its performance, AM allows

for customization by tailoring the unit cells to demonstrate specific bulk properties.

Figure 1.13 One method of generating a functionally graded lattice is by computing the stress within

the part, plotting the stress in gray scale, and then dithering the image to generate a density map.

Next, small stiff unit cells are placed in dense regions, with high stress, and large unit cells are placed

in less dense regions, adapted from [229].

Page 61: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

32

Figure 1.14 Functionally graded lattice structure populated using a relative density map based on CT

imaging, adapted from [178]

Page 62: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

33

Figure 1.15 Lattice generation and sizing method using stress results from a finite element analysis,

adapted from [234]

Problem Definition

Base-Lattice

Topology Generation

Solid-Body

Analysis

Page 63: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

34

In addition to these methods being researched, there is a variety of commercial software

that offer lattice structure population within a part as well as some capabilities to generate custom

lattice structures. Among the prominent commercial software is nTopology [235], Paramatters

[236], Netfabb Ultimate [126], 3DXpert [237], Discovery Spaceclaim [238], 3D CASTS

(Computer Aided System for Tissue Scaffolds) [170,175,239–241], or software developed by

Vongbunyong and Kara [242,243]. While all of this software has been developed to ease the

workflow for lattice structures in AM components, there is limited ability within the software to

provide feedback for selecting the proper unit cell topology for the application. For example, if a

company wants to populate a lattice inside their component, they can do so using Netfabb

Ultimate [126]; there are even some limited options for optimizing their lattice structure based on

the loading conditions. However, there is no way to know which of the 19 lattice structure options

is ideal for their application. Any of the 19 lattices could be populated within the component and

subsequently optimized without a knowledge of which lattices are manufacturable or provide the

desired bulk mechanical properties. In addition, most software only offers pre-generated unit cell

types, limiting the options and variations that are possible with AM.

Of these generation and population methods in research or commercial software, there is

a lack of reasoning or a systematic approach for why a particular unit cell is chosen. Essentially

any unit cell could be populated based on the density mapping or sized based on the finite

element results and there is no clear logic provided as to why a unit cell is chosen over another.

While there is data available for a variety of lattice structures, there is a need for design guides

and unit cell generation tools that may be used to understand the current unit cell structures or

customize the lattice structure for the application. AM has the capability to manufacture

multifunctional unit cells and lattice structures but there are currently limited options for

designing or optimizing unit cells based on application-specific requirements [116].

Page 64: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

35

1.1.5.1 DfAM of Lattice Structures

For PBF, there are several challenges with printing lattice structures including four

discussed by Nazir et al. [244] in their review of lattice structures fabricated with AM: (1) support

material removal, (2) powder removal, (3) large file size and (4) high surface roughness. Similar

to large components, build angle and the requirement to anchor low build angles and overhangs to

the build plate remain a concern for lattice structures. Due to the intricate complexity and feature

size of lattice structures, support material removal quickly becomes infeasible or impossible. For

powder removal, Koizumi et al. [245] specifically mentions the challenge of removing powder

for low porosity lattice structures. Low porosity lattice structures can trap powder due to minor

manufacturing variations that close off connected pores. The speed of post-processing can also be

affected by the time required to evacuate all the loose powder through a complex maze of

interconnected pores. Regarding file size, the most common file format for 3D-printing is the

Stereolithography file (STL). An STL is simply a triangulated surface mesh of the component.

Due to the intricate surface complexity of a lattice structure, an STL is ill-suited to be able to

accurately represent a component in a compact format, due to a file size that quickly explodes,

even for small components. As such, there are a variety of research areas being explored to

improve modeling and STL (or mesh) generation for lattices compared to traditional software

[246–249]. In practical cases, file size for lattice structures can be reduced by changing the cross

section of the struts from circular to hexagonal or octagonal to make the tessellation simpler.

Depending on the diameter of the lattice struts, the file size can be significantly reduced without

impact on the structure because the simplification of circular to hexagonal or octagonal can

become blurred out by the minimum feature size of the process. In addition, some commercial

software and production systems are moving away from the traditional STL file format due to

issues with file size and an inaccurate tessellated representation of the object. While all AM

Page 65: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

36

components are subject to poor surface roughness, the effects are compounded by the fact that the

feature sizes of lattice structures are much closer that of the surface roughness. Compared with

macroscale topology, surface variations on lattice structures have a larger impact on the expected

properties and performance. In addition to the four limitations covered by Nazir et al. [244],

Maconachie et al. [250] and others [194,223] also address the staircase effect which can cause

significant changes in cross section for thin lattice features, negatively impacting the mechanical

properties from the intended design. Others have reported on bending or deformation of intricate

or delicate structures during the printing process [251,252].

1.1.6 Design Optimization

For many years various algorithms have been proposed to optimize a design, especially

valuable for a manufacturing process with substantial design freedom. With the “free complexity”

allowed by AM, the ability to manufacture complex optimized shapes becomes a reality. While

AM remains an expensive process for large batch sizes, the geometric complexity of a one-off

part does not necessarily directly relate to the cost. This fact allows for geometrically complex

and customizable parts to be reasonably manufactured. With this reality, areas such as

customization and design optimization have seen an explosion of interest from the research

community. Optimization algorithms that have been researched for many years to maximize

performance in theoretical components. With AM, the cost and ability to manufacture these

optimized components is more reasonable than traditional subtractive manufacturing processes.

This section addresses two areas of optimization: common optimization algorithms and topology

optimization.

Page 66: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

37

1.1.6.1 Optimization Algorithms

In design optimization, numerous algorithms have been proposed including a variety of

bio-inspired algorithms, mathematical based algorithms such as gradient search, and other

physics based methods [253–255]. Bio-inspired algorithms include several different methods

based on principles in natural such as natural selection or evolutionary algorithms, swarming of

birds, or insects. Gradient search or gradient descent is a common method with several variations

for calculation or estimation of the gradient. Other physics-based models mimic physical

behavior or laws of physics. For all optimization algorithms and throughout the dissertation, there

are several terms used to describe the optimization. The objective function refers to a measure of

the goal of the optimization, sometimes called the cost function. The objective function, by

convention, is typically minimized to find the best solution. The design variables are the set of

values which may be changed to create different possible solutions. For each variable there is

usually an upper and lower bound which define feasible solution space. In addition to upper and

lower bounds, additional constraints impose restrictions on the solution space that relate two or

more optimization variables. A penalty function is similar to a constraint but is typically applied

after evaluation of the objective function. In addition, a constraint completely restricts certain

designs, whereas penalty functions have the flexibility of discouraging certain designs by

negatively impacting the objective function, without necessarily removing the design from the

solution space.

Evolutionary algorithms are based on several principles including reproduction, mutation,

gene crossover, and natural selection. Of evolutionary algorithms, the genetic algorithm is the

most common [256]. In general, a population of possible solutions are formed randomly

throughout the solution space, this set of individual solutions is called the population in the first

generation. Each individual in the population is evaluated based on the objective function(s), also

Page 67: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

38

known as the fitness function(s). The high performing designs according to the fitness function(s)

are selected as parents for the next generation. The parents, or dominant designs, are used to

“reproduce” and create children designs that are a mixture of these dominant designs. This

reproduction or mixture process of the parents is often referred to as crossover, consistent with

the chromosome crossover that happens in genetics. These children form the subsequent

generation, creating a new population of designs. In reality, the subsequent generations are mixes

of the dominant designs, essentially attempting to breed better individuals within each generation.

In addition to crossover, another process called mutation is also implemented for a small

percentage of each population. It allows some portion of the mixture for a child to be randomly

generated, creating a mutation for the child. The mutation process is used to help expand the

search of the design space by randomly searching other areas of the solution space. The algorithm

progresses until the convergence criterion is met, typically based on how much the population is

changing with each generation. A general diagram of a genetic algorithm is shown in Figure 1.16.

There are several genetic algorithms and variations that have been proposed, but one of the most

common algorithms is the Non-dominated Sorting Genetic Algorithm (NSGA-II) [257] for multi-

objective optimization problems, discussed in Section 2.3.2. In addition to NSGA-II, Borg is

another multi-objective evolutionary algorithm which incorporates a variety of search methods

and identification of the top performing designs [258]. Additional information on Borg is

available in Section 3.2.2.2.

Page 68: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

39

Figure 1.16 Diagram of genetic algorithm optimization process

Gradient search algorithms numerically estimate or analytically calculate the gradients of

the objective function to find the minimum or maximum of the objective function. A cartoon

image description of a gradient search algorithm for maximization of an objective is shown in

Figure 1.17. In the cartoon, it depicts a blindfolded person trying to find the top of the hill. The

height on the hill is equivalent to the objective function. In this case, the goal is to maximize the

objective function or find the highest point on the hill. The position of the blindfolded person is

the optimization variable. The fences define constraints or bounds on the feasible solution space.

As the blindfolded person takes a step, they can feel whether they are moving up or down the hill

and adjust accordingly. Eventually the person will reach the highest point on the hill at the red

flag. Just as described the cartoon, the gradient search method follows a similar process. By

taking a small step in the solution space, the gradient of the objective function or change in

objective function is calculated determining which direction to move. There are several variations

Page 69: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

40

of the gradient search algorithm based on different methods to calculate or estimate the gradient

and acceleration methods for faster convergence [259–261].

Figure 1.17 Cartoon graphic used to describe maximization of an objective for gradient-based

algorithms, adapted from [254]

Referring again to Figure 1.17 to compare genetic and gradient-based algorithms, a

genetic algorithm would start with a random population of people in the design space. The

individuals highest on the hill would be used to create children in attempts to find a higher

position on the hill than their current positions. The process would be repeated with an occasional

random mutation to try to search other areas until the population is no longer finding higher

locations on the hill. As a quick comparison, initially the genetic algorithm is able to get a sense

for the design space, assuming the space is well-sampled by the initial random population. Using

the parents of the population the objective value, or average height on the hill for a population

should increase quickly but it is not guaranteed that the crossover of parents will generate lower

objective values than the parents. It may take many generations of gradual improvements to

progressively increase the objective function. In contrast, the gradient-based algorithm begins

Page 70: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

41

with only a single objective function evaluation and should be able to continuously improve the

objective function unless there is difficulty with calculating or estimating the gradient.

Genetic and gradient-based optimization algorithms have been compared for many

different applications from which a general comparison can be formed between the two

optimization methods. In general, Salomon [262] compared the two methods and highlighted the

benefit of the population that exists for genetic algorithms. The population is argued to provide a

parallel search method as opposed to the gradient method which progresses from a single

solution. However, arguably, Salomon [262] points out that the gradient-based method does

search multiple directions to find the steepest gradient. The parallel search of a population in

genetic algorithms has also has shown a greater tendency to find the global minimum than

gradient-based algorithms [262]. The small step size of gradient-based algorithms, required to

maintain stability, makes it difficult to not become stuck in local minimums. Zingg et al. [263]

describes the tradeoff in cost between genetic and gradient-based algorithms. The high cost of the

genetic algorithm is based on the number of calculations which must be performed. Zingg et al.

[263] found that the genetic algorithm required between 6 and 200 times more function

evaluations than a gradient-based algorithm. The high cost of the gradient-based approach is the

requirement that the gradient must be computed. This can be challenging for non-smooth

objective functions or where the gradient cannot easily be defined [262]. In cases where the

gradient is readily available, Zingg [263] describes the primary advantage of the gradient-based

approach being rapid convergence. In addition, Zingg [263] points out that the convergence

criteria for gradient-based approaches is straightforward. As the gradient decreases by orders of

magnitude, one can be sure that at least a local minimum has been found [263]. In contrast, for

genetic algorithms they are able to handle non-smooth objective functions because a

mathematical relation between the optimization variables and objective function is not required

[264]. Genetic algorithm convergence is generally based on how much change exists in the

Page 71: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

42

population but it is not as straightforward. Because genetic algorithms often excel where gradient-

based algorithms do not, and vice versa, performance often varies based on application [265].

Many researchers have also attempted to combine them to get the best of both algorithms

[262,266,267].

Optimization algorithms may be applied to a variety of different problems using one or

more objectives. In the case of multi-objective optimization algorithms, there is generally no

single best solution but instead there is often a tradeoff between objectives because the objective

functions may conflict with one another. The multi-objective nature results in a set of solutions,

each being optimal depending on how important each objective function is. This set of optimal

solutions is referred to as the Pareto front and is defined by the set of solutions which are non-

dominated by any other solution. A non-dominated solution is a solution for which no other

solution is better in all objective functions. Multi-objective optimization problems present a

unique advantage because all solutions along the final Pareto front are optimal. On the contrary,

weighting factors or ratios are required when multiple objectives are combined into a single

objective. To test different weights or ratios, the optimization problem has to be run multiple

times. A multi-objective optimization enables the designer to view a set of Pareto optimal

solutions and select one based on the relative importance or limits of each objective function.

1.1.6.2 Topology Optimization

Topology Optimization is a method to optimize material placement throughout a design

space. Originally proposed in 1988 by Bendsøe and Kikuchi [268], it was based on discretizing a

design space into small elements and using optimization techniques to determine which elements

should contain materials and which should be void. After determining the size of the design space

and applying boundary and loading conditions, the design space is discretized where each

Page 72: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

43

element contains a “density.” Using a homogenization technique to determine the bulk properties

of the overall design space based on the density of individual elements, an optimized topology

can be generated. One of the challenges with the initial homogenization approach was the slow

convergence and difficulty in finding a clear boundary around the geometry. Because the density

value for each element can range from 0-1, 0 being void and 1 being full material, there was an

inherent gray-area where it was unclear whether material should be present or not. Later, the

Solid Isotropic Material with Penalization (SIMP) method was developed to improve

convergence and clearly define the boundaries of the solution [269,270]. The SIMP method uses

a power-law approach to drive the element densities to away from the gray area range in the

middle and force them closer to 0 or 1. The power-law approach acts as a penalization factor for

element densities in the mid-range. An example of the effect of varying the penalization

parameter for a minimization of compliance benchmark is shown in Figure 1.18. Traditionally

topology optimization was based on minimization of compliance and a material volume

constraint, with the intent of creating stiff and light-weight structures. For many years, topology

optimization has been developed, now using a variety of approaches and applied to complex

situations such as compliant mechanisms [271–276], multiple load cases [277], heat transfer

[118,278], multiple materials [13,279], and constraints for different manufacturing processes

[280]. A review of topology optimization methods and applications is presented by Bendsøe and

Sigmund [281] and Sigmund and Maute [282], and a review of topology optimization in

commercial software is provided by Reddy et al. [283].

Page 73: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

44

Figure 1.18 Minimization of compliance benchmark optimized with the SIMP method. Varying the

power-law penalization factor, p, reduces the amount of gray material in the solution, clearly

defining the boundaries of the topology. Images adapted from [270].

In the case of compliant mechanism design or tailored compliance, one method of design

synthesis is topology optimization. Often in topology optimization of structures, the objective of

the optimization is minimization of compliance. On the other hand, for tailored compliance or

compliant mechanisms, compliance is desired, requiring new objectives for optimization. Often

the objective is a coupling between deflections at an input point and output point. To determine

the input/output relationship, the standard minimization of compliance problem is altered to

account for motion or force at the output point. Ananthasuresh and Frecker [284] describe

compliant mechanism synthesis objective functions based on mutual strain energy and strain

energy. The mutual strain energy is a measure of the output displacement and the strain energy is

a measure of the inverse of stiffness. These are conflicting objectives that must be balanced

through optimization. For example, the desired output force or displacement must be balanced

with a structure that stiff enough to handle the external loads. Ananthasuresh and Frecker [284]

describe two compliant mechanism synthesis approaches using topology optimization: ground

Page 74: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

45

structure parameterization and a continuous material density parameterization. The ground

structure parameterization method is discussed in the following paragraph and the continuous

material density parameterization was described previously, though the objective functions for the

optimization must be reconsidered as described by Ananthasuresh and Frecker [284]. Many

examples of topology optimization and compliant mechanisms exist in the literature, especially

for the classic example of a compliant displacement inverter, Figure 1.19, and compliant gripper,

Figure 1.20.

Figure 1.19 Boundary and Loading conditions for a compliant inverter as well as the evolution of the

topology during convergence, adapted from [285]

Page 75: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

46

Figure 1.20 A variety of compliant gripper topologies generated by varying optimization parameters,

adapted from [285]

One class of topology optimization methods is a called ground structure topology

optimization (GSTO). GSTO is known by many other names such as, structural optimization,

structural topology optimization, truss optimization, truss topology optimization, and ground

structure optimization. In contrast to traditional topology optimization, GSTO methods are a

discrete approximation of the design space by using truss, beam, or frame elements connected to

nodes throughout the design space. While in traditional topology optimization the design space is

discretized using continuum elements, the GSTO method approximates the design space as a set

of nodes distributed throughout and truss elements connecting the nodes; see Figure 1.21.

Assigning a cross-sectional area value to each member, the domain may be optimized with a

volume constraint for a given set of boundary and loading conditions. The earliest references to

GSTO dates back to the 1960’s with work by Dorn et al. [286]. The methodology and formulation

Page 76: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

47

have been described by Bendsoe and Sigmund [281] and Ohsaki [287] in their textbooks on

topology optimization. GSTO has since advanced to tackle a number of different challenges

including multi-material design [288], reliability [289], arbitrary 2D and 3D domains [290,291],

adding and removing nodes or elements [292], buckling and overlap [293], and other AM

considerations [294–296].

Page 77: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

48

Figure 1.21 An example comparison of the continuum and ground structure approach for topology

optimization

Continuum Approach Ground Structure

Approach

Problem Definition

Page 78: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

49

1.1.6.3 Unit Cell Topology Optimization

In 1995, Sigmund researched a method for using a GSTO to generate a representative

base cell, or unit cell topology, for mesostructure design [297,298]. Using homogenization, the

unit cell topology was optimized for tailored compliance, or desired bulk mechanical properties.

Though it was mathematically possible to solve for the unit cell topology, the ability to

manufacture complex mesostructures was severely limited. Since then, multiple uses of this

approach have been used such as the work by Guth et al. [299,300], Messner [301], Neves et al.

[302], Liu et al. [303], and Huang et al. [304]. In 2012, Guth et al. [299] reported work very

similar to that of Sigmund, showing several 2D examples of unit cells tailored to specific bulk

properties. In 2015, Guth et al. [300] then reported 3D unit cell examples while considering

matching thermal and mechanical properties while enforcing isotropy constraints. Messner [301]

reported a broader use of the approach by not limiting the bounding volume of the unit cell to a

predetermined size and shape. Neves et al. [302] reported on the use of a penalization factor for

localized buckling in slender beams to increase the stability of the unit cell. Liu et al. [303] and

Huang et al. [304] both used the approach to develop cells of minimum compliance. A variety of

other works have also focused on homogenization and comparison of different techniques [305–

309].

In addition to the ground structure and homogenization method proposed by Sigmund, a

variety of other unit cell optimization methods have been proposed. Chen et al. [133], Xia et al.

[310], and Neves et al. [311] describe the use of continuum-based topology optimization to

design novel lattice structures with tailored properties. Ground structure approaches have also

been used to explore unit cell optimization with uncertainty and robust design considerations

[118,312]. Watts et al. [313] and Watts and Tortorelli [314] have published methods for reducing

computational cost of lattice structure design through surrogate models and geometric projection

Page 79: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

50

methods. Geometric projection relies on geometric primitives that are projected onto a continuum

mesh for finite element analysis [315]. It has been applied through various way including simple

cylindrical primitives [316], plates [317], or supershapes [318]. Other methods have also

proposed hybrid approaches that combine topology optimization of the overall geometry with

sizing the lattice structure [319–324]. While there are numerous research groups exploring

methods for unit cell optimization or metamaterial design, the direct incorporation of AM

constraints into the optimization of unit cells has not been addressed.

1.1.7 Topology Optimization and Additive Manufacturing

With the growing interest in AM discussed previously, there has been a renewed interest

in topology optimization. Topology optimization methods often produce complex geometries that

are organic in nature and tend to be difficult, if not impossible, to manufacture with traditional

methods. While there are many complications to overcome, it is often possible to directly print

topology optimized solutions because the part is manufactured directly from a computer model.

One challenge with topology optimization for AM, is how to include manufacturing constraints

and consideration, such as build time and cost [325]. Without manufacturing constraints, the

solution generated does not account for process capabilities or limitations.

In 2016, Liu et al. [280] published a review of how manufacturing constraints for

machining, injection molding, and casting are implemented in topology optimization and then

highlighted topology optimization for AM as a promising future research direction. Then, in

2018, Liu et al. [295] provided a thorough review specifically for topology optimization for AM.

In the review, he highlights the importance of support material considerations for topology

optimization and the approaches that various research groups have taken. Langelaar [326,327] has

created an AM filter for post processing topology optimized designs into an as-printed part,

Page 80: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

51

including a 2D and 3D implementation. In addition, Langelaar [328] proposed a cost function to

understand the tradeoff between restricting the topology optimization to be overhang-free and

unrestrained topology optimization; see Figure 1.22. Brackett et al. [329] proposed an idea for

using a linear approximation of overhangs to determine the build angle and penalize designs with

low build angles. Gaynor and Guest [330] used an integrated approach to include the critical

overhang angle within the topology optimization fully self-supporting components. Another

option that Liu et al. [295] highlights for reducing support structures is support slimming. Liu et

al. [331] proposed a method for detecting and restricting closed voids to avoid trapped powder.

While various research groups continue to explore methods for incorporating DfAM into

topology optimization, there is a lack of research which explore DfAM constraints on unit cell

topology optimization for designing lattice structures.

Figure 1.22 Potential tradeoffs exist in the ability to find design that consider the amount of support

material used while simultaneously considering the performance of the component, adapted from

[328]

1.1.7.1 Combining Lattice Structure DfAM and Topology Optimization

The second area of topology optimization for AM addressed by Liu et al. [295] is porous

infill and mesoscale design. Their work highlights lattice optimization or complex mesoscale

design as an important research area of topology optimization and AM. Due to the length scale,

topology optimization should produce fully self-supporting designs because of the difficulty

required to remove supports, especially for metal AM. Regarding mesoscale design, researchers

Page 81: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

52

have produced topology optimized structures for tissue engineering scaffolds [168,169], auxetic

structures [332], other tailored structures for biomedical applications [245,333]. Though the area

of topology optimization for tailored properties has been an area of research for many years, as

described in Section 1.1.6.3, there is a lack of research for incorporating DfAM. In particular,

DfAM and unit cell topology optimization using a ground structure approach for tailored lattice

properties has not been addressed. With much of the focus for topology optimization and AM

centered on making stiff and light-weight structures, there is still much to be explored for tailored

compliance or multifunctional structures. In particular, the free complexity of AM that allows for

mesostructural design may be blended with GSTO synthesis for a variety of objective functions

including tailored bulk properties, thermal conductance, and AM consideration such as powder

removability.

There are several examples of topology optimization used to design novel metamaterials

with unique mechanical properties [133,197,304], biomedical purposes [163,245], or other multi-

functional lattices [118]. However, many of these approaches give little regard to incorporation of

the manufacturing process constraints into the optimization. In 2018, Tamburrino et al. [334]

review the design process for lattice structures fabricated with AM. As a result of their review,

they call for further research in five areas:

1) strategies to predict mismatch between as-designed versus as-manufactured,

2) optimization strategies for material distribution for enhanced and customized

properties allowed by AM,

3) optimized models that reduce computational effort,

4) integrated design tools for a multi-disciplinary approach, simultaneously taking into

account technological, geometric, and technical requirements, and

5) development of novel functionalities or properties that go beyond simple

lightweighting.

Roadmaps and reviews of the current status and future roadmaps have come to similar

conclusions [122,295,335–337]. Liu et al. [295] further supports this call for improved

incorporation and consideration of DfAM into the optimization. Plocher et al. [336] refer to

Page 82: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

53

DfAM constraints in optimization as a, “bottleneck for a streamlined design procedure”. In

summary, there is a need for multi-objective optimization methods for lattice structure design that

incorporate DfAM constraints and variations that may occur during printing. This dissertation

presents a systematic approach for the optimization of unit cell topology that directly incorporates

DfAM considerations and demonstrates it with multiple case studies.

1.2 Research Objectives and Tasks

Recent advances in the capabilities of model simulation and manufacturing technology

are paving the way for systematic design optimization methods. These methods enable designers

to explore organic and nonconventional designs that were previously not possible to analyze or

manufacture. Through design optimization, these nonconventional designs may be used to

maximize part performance in a variety of applications such as medical, aerospace, and defense

industries.

With the increased need for design optimization tools for applications across multiple

industries, the following research objectives and tasks are outlined below for the development of

two systematic design optimization approaches.

1. Develop systematic optimization approach for compliant deployable radiofrequency

ablation electrodes

1.1. Design deployable electrode for endoscopic radiofrequency ablation

1.2. Develop finite element model to simulate radiofrequency ablation in soft tissue

1.3. Couple finite element with optimization approach for design of electrode based on

tumor geometry

1.4. Evaluate the deployable electrode design

1.4.1. Assess deployment of early prototypes in tissue phantom

1.4.2. Validate optimized electrode design through ex-vivo experimentation

2. Develop three-dimensional ground structure topology optimization for unit cell DfAM

2.1. Develop 3D-parametric ground structure model

2.2. Establish systematic optimization method for unit cell generation

Page 83: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

54

2.2.1. Generate set of potential objective functions and constraints

2.2.2. Quantify manufacturing considerations for metal AM

2.3. Validate optimization with benchmark problems

3. Explore multi-functional lattices through case study on unit cell generation for thermal

conductance and minimum strain energy

3.1. Define constraints and process limitations

3.2. Generate optimal lattice structure with and without DfAM

3.3. Fabricate lattice structure with DfAM considerations

3.4. Evaluate solutions and revise ground structure topology optimization

4. Explore homogenized lattices through case study on generation of unit cells with tailored

compliance

4.1. Define constraints and process limitations

4.2. Generate optimal lattice structure with tailored properties

4.3. Fabricate and test to determine mechanical properties

4.4. Characterize fabricated samples with mechanical testing

4.4.1. Compare boundary conditions of homogenization with experimental

characterization

5. Develop mechanical property database for metal lattice structures from current literature

containing analytical, finite element, and experimental data

5.1. Perform literature search for metal lattice structure data

5.2. Collect data from selected papers

5.3. Create user interface for generating mechanical property charts of collected data

1.3 Dissertation Contents and Structure

The remainder of the dissertation is structured in the following format. Chapter 2

addresses work related to objective 1. The design of a deployable compliant endoscopic RFA

electrode is detailed. Next, a finite element model to predict the treatment zone for

radiofrequency ablation is explained. Then, a systematic optimization approach for shape

matching of an RFA treatment zone with electrode shape is explained. The optimization approach

is demonstrated in treatment of a pancreatic tumor using two variations of the proposed compliant

electrode. Finally, the deployment testing of the electrode prototypes and validation of the

computational model with a thermochromic tissue phantom is described.

Chapter 3 describes a systematic optimization approach for unit cell generation using a

3D GSTO, addressing objective 2 and its tasks. The GSTO approach is called Additive Lattice

Page 84: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

55

Topology Optimization (ALTO). DfAM optimization objective functions, constraints, and

penalty functions were applied to improve manufacturability for AM. Next, a set of potential

objective functions, constraints, and penalties for a variety of applications is presented. Finally,

comparison of the ALTO results for optimization benchmark test problems validate the approach.

Chapters 4 and 5 address the tasks related to objectives 3 and 4, detailing the case-studies

included in this work. The case studies include 2D and 3D generation of novel unit cell

topologies for (1) thermal conductance and mechanical stiffness, (2) tailoring the mechanical

properties of a unit cell while maintaining a constant volume fraction, and (3) improving powder

removability while maintaining mechanical properties. Each case study includes description of

the purpose, optimization parameters, and results. From each case study, the outcomes and insight

gained are discussed, including adjustments and acknowledgement of limitations of the ALTO

approach.

In Chapter 6, the development of Lattice Unit-cell Characterization Interface for

Engineers (LUCIE) is presented. LUCIE combines published literature surrounding analytical,

finite element, and experimental analysis of lattice structures into MATLAB-based GUI for

comparison of mechanical properties of unit cell topologies. The selection of a unit cell topology

is available in commercial software; however, the selection is essentially blind because there is no

information regarding unit cell selection for the intended application. Though still in its infancy,

the future of topology optimization and AM, such as ALTO, would allow for complete design

and optimization of a customized lattice structure within a component. LUCIE offers a bridge

between rigorous optimization and blind selection in commercial software by providing a

centralized database for mechanical characterization of unit cell topologies.

Finally, Chapter 7 is a brief overview of the entire work. The overview includes major

outcomes and conclusions regarding each objective and task. To conclude, the research

contributions and recommendations for future research directions are addressed.

Page 85: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

56

Chapter 2 Optimization and Experimental Validation of an EUS-RFA

Electrode

The focus of this chapter is on the development of a systematic optimization approach for

a deployable EUS-RFA compliant electrode. It includes a brief summary of the relevant

background and motivation as well as descriptions of the proposed electrode design, a finite

element model to predict the RFA in tissue, the optimization approach, results to date, and

experimental validation of the electrode design. This entire work was performed in collaboration

with Dr. Moyer from Hershey Medical Center. The following objectives and sub tasks are

addressed in this chapter:

1. Develop systematic optimization approach for compliant deployable radiofrequency

ablation electrodes

1.1. Design deployable electrode for endoscopic radiofrequency ablation

1.2. Develop finite element model to simulate radiofrequency ablation in soft tissue

1.3. Couple finite element with optimization approach for design of electrode based on

tumor geometry

1.4. Evaluate the deployable electrode design

1.4.1. Assess deployment of early prototypes in tissue phantom

1.4.2. Validate optimized electrode design through ex-vivo experimentation

2.1 Background and Motivation

Radiofrequency Ablation (RFA) is an increasingly used, minimally invasive, cancer

treatment modality for patients who are unwilling or unable to undergo a surgical resection of

abdominal tumors. RFA is performed by heating the tissue with an electrode inserted into the

tumor. The shape of the treatment zone is primarily determined by the shape the electrode and the

thermal and electrical properties of the tissue. While it has shown great promise in treatment, it is

limited by tumor location, tumor geometry, and tracking the treatment zone. Endoscopic

Page 86: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

57

ultrasound-guided radiofrequency ablation (EUS-RFA) offers access to additional regions of the

abdomen that were previously inaccessible through traditional percutaneous RFA. Current EUS-

RFA electrodes are straight needle-shaped, generating an elliptical ablation zone, and offer

limited ability to treat the entire tumor; see Figure 2.1. This requires multiple insertions and

treatments to destroy a single tumor. To understand RFA, numerous finite element models have

developed; see Section 1.1.3 for additional information. While computational models have been

developed for understanding the effect of tissue properties, heat sink effects, placement

optimization, and commercial electrodes, there have been no attempts to use optimization to

determine the design of an RFA electrode for a specific tumor shape. There is a need to develop a

systematic optimization approach for generating RFA electrode designs to effectively treat tumor

geometries. In this chapter, a systematic optimization approach for a deployable EUS-RFA

electrode is applied specifically to pancreatic cancer, however, it is important to note that the

approach is not limited to pancreatic cancer. The fundamental approach and analysis may be

applied to cancers throughout the abdomen and mediastinum.

Figure 2.1 For straight electrodes, multiple insertions are often required to overlap a series of

ablation zones to destroy tumors. The image above is for an example of a percutaneous approach,

but, for EUS-RFA, the ability to insert multiple electrodes is limited creating the need for specialized

electrodes, adapted from [51].

Page 87: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

58

2.2 Deployable Compliant EUS-RFA Electrode Design

The compliant EUS-RFA electrode design was inspired by the tail feathers of a peacock,

which begin as a straight bundle and then expand out into a large circular display. An earlier

version of the design was proposed by Hanks et al. [338]. The peacock-inspired electrode was

designed to be introduced into the tumor endoscopically, through a hollow needle. As shown by

Figure 2.2, the endoscopic needle, or sheath for the electrode, is advanced through the working

channel of an endoscope up to the periphery of the tumor. Then, the electrode is delivered

through the needle and deployed into the tumor. The design of the electrode is such that it may be

retracted into the endoscopic needle and, if necessary, redeployed in another location.

Figure 2.2 The electrode is deployed into the tumor through an endoscopic needle, acting as a sheath,

whose tip has been positioned at the periphery of the tumor, adapted from [339]

Two variations of a generic or base design for the optimization, shown in Figure 2.3, are

based on a circular array of outer tines and a single central tine which are connected at the

exposed electrode base. It is shown in both the stowed (Figure 2.3A,D) and deployed (Figure

2.3B,E) configurations. The principle of the design was that the natural state of the device was

open. The outer tines were compressed around the central tine and stowed within a rigid sheath,

such as an endoscopic needle. When the electrode was extended out of the rigid sheath, the

Page 88: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

59

electrode tines would deploy and cut paths radially outward into the tumor facilitated by their

beveled tips. The deployment of the electrode tines broadens the ablation zone as compared to a

traditional straight needle electrode. The deployable electrode design was parameterized in terms

of the outer diameter in the stowed configuration (𝑑𝑒), length of the center tine (𝑙𝑐), number of

outer tines (𝑛𝑡), length of each outer tine (𝑙𝑜𝑖 , 𝑖 = 1,… , 𝑛𝑡), radius of curvature of each outer tine

(𝑟𝑜𝑖, 𝑖 = 1,… , 𝑛𝑡), and length of the exposed electrode base (𝑙𝑏) which connected the outer tines to

the center tine. Figure 2.3C shows the electrode design parameters 𝑑𝑒, 𝑙𝑐, 𝑙𝑜𝑖 , 𝑟𝑜

𝑖, and 𝑙𝑏 for the 𝑖-

th outer tine of an 8 outer tine example (𝑛𝑡 = 8). Figure 2.3F shows the design parameters

associated with a 12 outer tine variation (𝑛𝑡 = 12) of the electrode which is used as a second case

for optimization. The outer tines were designed to have constant curvature because studies of

flexible needles showed that during insertion, the curvature of the needle path was essentially

constant [340–342].

Page 89: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

60

Figure 2.3 Two variations of the proposed electrode design are pictured in the stowed (A,D) and

deployed (B,E) configurations. The design parameters for each are shown (C,F), adapted from [339]

Page 90: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

61

2.3 Systematic Optimization Approach

2.3.1 Ablation Zone Prediction with a Finite Element Model

The Electric Currents and Bioheat Transfer modules of COMSOL Multiphysics

Modeling Software Version 5.2a [343] were used to develop the thermal ablation model (TAM).

These modules coupled the thermal and electrical components of the simulation of RFA in

biological tissue through the Multiphysics module. The finite element analysis was transient and

simulated an ablation procedure to obtain the geometry of the ablation zone generated by an

electrode. During the simulation, the electrode design was held fixed, as defined by design

variables shown in Figure 2.3. The TAM used similar underlying assumptions and boundary

conditions as the finite element models by Chang [61] and Tungjitkusolmun et al. [67].

A cylindrical block with a radius of 𝑟𝑡 and height ℎ𝑡 was used to represent the tissue and

must be large enough to avoid boundary interactions with the elevated temperatures generated by

the ablation. An insulated shaft, that would deliver the power to the electrode in the actual

procedure, was represented by a cylinder of radius 𝑟𝑏 and height ℎ𝑏. It was placed concentrically

inside the cylinder extending from the bottom of the tissue up to the bottom of the exposed

electrode base. The compliant electrode design was generated in COMSOL and fixed to the

insulated shaft, centered within the tissue block. To reduce the computational size of the TAM, a

simplified design without the beveled tips was generated for the ablation modeling. RFA tissue

heating was caused by the concentrated electric field near the electrode and thus heat transfer in

the tissue remained unaffected by the simplified TAM.

The fundamental equations used to model the thermal ablation are shown in Eqn. 2.1 and

Eqn. 2.2. (the bioheat equation).

∇ ∙ (−𝜎∇𝑉 + 𝑱𝑒) = 𝑄𝑗 (2.1)

Page 91: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

62

𝜌𝐶𝜕𝑇

𝜕𝑡+ 𝜌𝐶𝒗 ∙ ∇𝑇 + ∇ ∙ (−𝑘∇𝑇) = 𝜌𝑏𝐶𝑏𝜔𝑏(𝑇𝑏 − 𝑇) + 𝑄𝑚𝑒𝑡 + 𝑄𝑒𝑥𝑡 (2.2)

In Eqn. 2.1, 𝜎(S/m) is the electrical conductivity, 𝑉(V) is the voltage potential, 𝑱𝑒 (A/m2)

is an externally generated current density and 𝑄𝑗(A/m3) is a current source. In Eqn. 2.2, 𝜌 (kg/m3)

is the tissue density, 𝐶 (J/(kg∙K)) is the tissue specific heat, T (K) is the temperature, 𝒗 (m/s) is

the velocity vector, 𝑘 (W/(m∙K)) is the tissue thermal conductivity, 𝜌𝑏 (kg/m3) is the blood

density, 𝐶𝑏 (J/(kg∙K)) is the blood specific heat, 𝜔𝑏 (1/s) is the blood perfusion rate, 𝑇𝑏 (K) is the

arterial blood temperature, 𝑄𝑚𝑒𝑡 (W/m3) is the heat source from metabolism and 𝑄𝑒𝑥𝑡 (W/m3) is

the heat source from the spatial heating.

The Electric Currents module available in COMSOL is based on Eqn. 2.1 In this TAM,

the electric field generated by the radiofrequency current in ablation was approximated by an

applied voltage, thus 𝑱𝑒 and 𝑄𝑗 are assumed to be zero. This simplifies Eqn. 2.1 to Eqn. 2.3

below.

−𝜵 ∙ (𝝈𝜵𝑽) = 𝟎 (2.3)

Electric boundary conditions of the TAM were an applied voltage to the active portion of

the electrode design (𝑉 = 𝑉0) and grounded external surfaces of the tissue block (𝑉 = 0). The

applied voltage which generated a direct current as opposed to the radiofrequency current used in

the actual procedure was a simplification, which was also used in the finite element model by

Chang [71]. The grounded external surfaces of the cylindrical tissue block were representative of

the large grounding pads used in the procedure to disperse the current.

The Bioheat Transfer module of COMSOL, which describes the heating that occurs in

biological tissue, is based on Eqn. 2.2, the bio-heat equation. With the velocity (𝒗) and metabolic

heat source (𝑄𝑚𝑒𝑡) set to zero, Eqn. 2.2 is simplified to Eqn. 2.4.

𝜌𝐶𝜕𝑇

𝜕𝑡+ ∇ ∙ (−𝑘∇𝑇) = 𝜌𝑏𝐶𝑏𝜔𝑏(𝑇𝑏 − 𝑇) + 𝑄𝑒𝑥𝑡 (2.4)

Page 92: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

63

The thermal boundary condition was specified as a constant temperature on the external

surfaces of the tissue block (𝑇 = 𝑇∞). Heat conduction was permitted through the active portion

of the electrode and exposed electrode base. Tetrahedral elements were chosen to mesh the tissue

and electrode. Element size limitations were imposed in three sections to accommodate the small

features of the electrode geometry without drastically increasing the computational size.

To simplify the TAM, quarter symmetry was used to reduce the computational time; see

Figure 2.4. On the symmetry planes used to divide the geometry, an insulation boundary

condition was specified. For the Electrical Currents module of COMSOL, the insulation boundary

requires that no current flows through the boundary. In other words, the electric potential was

assumed to be symmetric about the boundary. For the Bioheat Transfer module, an insulation

boundary condition requires that the heat flux through the boundary be zero, or that the

temperature was symmetric about the boundary. For a summary of the initial and boundary

conditions; see Table 2.1.

Page 93: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

64

Figure 2.4 General setup for the TAM. Due to the symmetry of the electrode, the TAM was reduced

to quarter symmetry to decrease computation time. The grounding pads regions are the top and

bottom quarter circles as well as the exterior wall of the cylinder (not shown) , adapted from [339].

Table 2.1 Initial and boundary conditions for the TAM

Initial Condition Boundary Condition

Tissue Electrical 𝑉 = 0 --

Thermal 𝑇 = 𝑇∞ --

Electrode Electrical 𝑉 = 𝑉0 𝑉 = 𝑉0

Thermal 𝑇 = 𝑇∞ --

Grounding Pad Electrical 𝑉 = 0 𝑉 = 0

Thermal 𝑇 = 𝑇∞ 𝑇 = 𝑇∞

Symmetry Plane Electrical 𝑉 = 0 𝒏 ∙ 𝑱 = 0

Thermal 𝑇 = 𝑇∞ −𝒏 ∙ 𝒒 = 0

The ablation zone was defined by the 60°C isothermal surface surrounding the electrode.

Time and temperature dependent models for cell death vary in literature; however, at 60°C,

literature agrees that cell death is essentially instantaneous [40,42,44–46]. Chang and Nguyen

[71] found that the 60°C isothermal surface was a good representation of the ablation zone by

comparing their finite element model with experimental data.

Page 94: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

65

2.3.2 Optimization using a Genetic Algorithm

The optimization of the electrode was performed using a multi-objective genetic

algorithm in MATLAB [344] and was an extension of our previous work [345]. A genetic

algorithm was chosen because of the increased likelihood of finding a global minimum compared

to gradient-based approaches, and its ability to converge to an optimal solution without the need

to calculate numerical or analytical gradients. The multi-objective genetic algorithm is part of the

Global Optimization Toolbox in MATLAB [346] and is a variant of the Non-dominated Sorting

Genetic Algorithm (NSGA-II) algorithm [257]. The purpose of the algorithm was to find the

Pareto-Optimal set of designs, or Pareto front, which were non-dominated by any other design. A

non-dominated design is a design for which no other design has better than the current design in

all objective functions.

The NSGA-II algorithm variant in MATLAB works by evaluating a series of design sets

using the TAM. Each design set, or generation, was created based on the results of previous

generations, gradually converging to the Pareto-Optimal set of designs. MATLAB created the

initial generation randomly distributed throughout the design space. The design space was

defined by the user-defined constraints and the upper and lower bounds of each design variable.

Next, the electrode geometry, defined by the design variables, was sent to COMSOL for

simulation and the objective function data from each simulation was tabulated. Following each

COMSOL simulation, MATLAB collected the data and evaluated the objective functions for the

design. After all the designs in the generation were simulated, they were ranked based on the

objective function (fitness) evaluations. The non-dominated designs were rank one, designs

dominated by one other design were rank two, and designs dominated by two designs were given

a rank of three, etc. After all the designs of the generation were ranked, a portion of the best-

ranked designs were carried over to the next generation and other designs were generated by

Page 95: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

66

creating crossovers and mutations according to MATLAB default settings. The crossovers and

mutations allow a genetic algorithm to sample a large design space and help avoid local minima.

The process was repeated for subsequent generations, gradually improving with each generation.

Multi-objective optimization algorithms present a unique advantage over single objective

algorithms because all designs along the final Pareto front are optimal. On the contrary, weighting

factors or ratios are required when multiple objectives are combined into a single objective. To

test different weights or ratios, the optimization problem would have to be run multiple times. A

multi-objective optimization enables the designer to view a set of Pareto optimal designs and

select a solution based on the relative importance or limits of each objective function for a

particular application. In the case of treating tumors, the multi-objective optimization allows the

designer to view the set of optimal solutions, consider the importance of each objective function,

and select the best design while considering other patient-specific factors such as surrounding

anatomy.

As the optimization progressed, the metric used to determine convergence was the

change in the distribution or spread change of the Pareto front. Once the Pareto front was

essentially stagnant, the algorithm has determined a set of optimal designs which represent the

best tradeoffs of multiple objective functions. The change in the distribution of the Pareto front

was computed as the average relative change in the spread of the optimal designs, where the

spread is a distance measurement of the distribution of designs along the Pareto front. The

algorithm was considered converged when the average relative spread change for the previous

𝑛𝑠𝑡𝑎𝑙𝑙 generations was less than 𝜂, the stall tolerance, and the current spread was less than the

average of the previous 5 generations. In other words, the distribution of the optimal designs has

had minimal changes for 𝑛𝑠𝑡𝑎𝑙𝑙 generations and the spread of designs was decreasing.

The purpose of the optimization was to match the size and shape of the ablation zone for

a particular electrode with the size and shape of target ablation zone. After all the variables and

Page 96: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

67

objective functions are described, a formal description of the optimization is given by Eqns. 2.5-

2.9. The design variables included in the optimization are 𝑙𝑐, 𝑛𝑡, 𝑙𝑜𝑖 , 𝑟0

𝑖 as defined previously. The

upper and lower bounds for each design variable, represented by Eqn. 2.5, define the design

space.

𝑀𝑖𝑛𝑖𝑚𝑖𝑧𝑒

(𝑓𝑜𝑣𝑒𝑟 (𝑡𝑖 , 𝑟𝑖(𝑙𝑐 , 𝑙𝑜𝑖 , 𝑟𝑜

𝑖)) , 𝑓𝑢𝑛𝑑𝑒𝑟 (𝑡𝑖, 𝑟𝑖(𝑙𝑐 , 𝑙𝑜𝑖 , 𝑟𝑜

𝑖)))

𝑆𝑢𝑏𝑗𝑒𝑐𝑡 𝑡𝑜

{

𝑙𝑐𝑙𝑜𝑤𝑒𝑟

𝑙𝑜𝑙𝑜𝑤𝑒𝑟

𝑟𝑜𝑙𝑜𝑤𝑒𝑟

} ≤ {

𝑙𝑐𝑙𝑜𝑖

𝑟𝑜𝑖

} ≤ {

𝑙𝑐𝑢𝑝𝑝𝑒𝑟

𝑙𝑜𝑢𝑝𝑝𝑒𝑟

𝑟𝑜𝑢𝑝𝑝𝑒𝑟

}

𝑓𝑜𝑟 𝑖 = 1,… , 𝑛𝑡

(2.5)

−𝑙𝑐 + 𝐶𝑙𝑖𝑙0𝑖 ≤ 0 (2.6)

𝑊ℎ𝑒𝑟𝑒,

𝑓𝑖 =𝑟𝑖 − 𝑡𝑖𝑡𝑖

(2.7)

𝑓𝑜𝑣𝑒𝑟 = ∑ 𝑓𝑖𝑛𝑜𝑣𝑒𝑟1

𝑛𝑜𝑣𝑒𝑟 𝑓𝑜𝑟 𝑓𝑖 > 0 (2.8)

𝑓𝑢𝑛𝑑𝑒𝑟 = ∑ 𝑓𝑖𝑛𝑢𝑛𝑑𝑒𝑟1

𝑛𝑢𝑛𝑑𝑒𝑟 𝑓𝑜𝑟 𝑓𝑖 < 0 (2.9)

Two objective functions were used to quantify the performance of each design, each

based on a spherical coordinate system centered in the ablation zone; see Figure 2.5. As

mentioned previously, the 60°C isothermal surface determines the shape of the ablation zone. The

radial distance was measured from the center of the ablation zone to this isothermal surface for

𝑛𝑔𝑝 points. The center of the ablation zone was determined taking the average of the highest and

lowest ablation zone coordinates along the Z-axis. Each radial distance measurement, 𝑟𝑖, was

normalized in terms of the target radial distance, 𝑡𝑖, resulting in 𝑓𝑖, a percentage over or under the

target value for a specific θ, φ location; see Eqn. 2.7. The resulting 𝑓𝑖 values were grouped by

whether they were greater or less than the target distance and then averaged into two objective

functions, 𝑓𝑜𝑣𝑒𝑟 and 𝑓𝑢𝑛𝑑𝑒𝑟; see Eqns. 2.8-2.9. The objective function 𝑓𝑜𝑣𝑒𝑟 focuses on

Page 97: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

68

minimizing the amount of tissue destroyed outside the target zone (healthy tissue beyond the

target zone). The objective function 𝑓𝑢𝑛𝑑𝑒𝑟 focuses on maximizing the amount of tissue destroyed

within the target zone. While these two objective functions were not always in direct conflict,

they result in a set of Pareto-optimal designs from which the designer can make a more informed

decision than a single objective optimization. The use of these objective functions allows for

consideration of optimal electrode designs ranging from no ablation outside the target region

(𝑓𝑜𝑣𝑒𝑟 = 0) to ablating the entire target region (𝑓𝑢𝑛𝑑𝑒𝑟 = 0) and a variety in between. The

constraint in Eqn. 2.6 allows the designer to control the length of the center tine relative to the

length of the outer tine through 𝐶𝑙𝑖.

Figure 2.5 Graphical description of the objective functions used for optimization, adapted from [339]

2.3.2.1 Objective Function Computation through Image Analysis

Several different objective functions were explored including sphericity calculations, a

ratio of the second moment of area through different cross sections, and finally the comparison of

a series of distance measurements for the target and simulated ablation zone. Ultimately, the

Page 98: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

69

distance measurements proved to be the most robust and may be easily applied to a variety of

target shapes beyond a sphere. These objective functions make the approach applicable to a

variety of tumor shapes and even patient specific optimization based on patient tumor geometry.

To compute the objective functions two methods have been utilized. The first methods

used probes internal to COMSOL to measure the distance to the edge of the ablation zone at

different locations. The second approach uses image analysis to compute the cross sections of the

ablation zone. Image analysis tools within MATLAB determine the center of the ablation zone

and measure the distance to the edge. The MATLAB code for performing the image analysis was

developed with support from Zackary Snow and is shown in Appendix A. In both approaches,

these distances were compared against the target shape to compute the objective functions.

Compared to the first approach, the second approach enables the designer to compare a greater

number of measurement locations between the simulated and target ablations, but ultimately the

results were similar. For arbitrary shapes, where many points would be necessary to define the

shape, the second approach may show better robustness than the first approach. As the results for

both approaches were similar, the results from the first method are described in the following

section.

The use of these objective functions resulted in a set of Pareto optimal electrode designs

for which the ablation zone geometry ranges from completely inside the target to variations that

fill more of the target zone. For the purposes of further analysis and comparison of the two

electrode variations shown in Figure 2.3, a single design was selected from the set of optimal

solutions. Without patient specific anatomy to consider in narrowing down the set of optimal

solutions, the distance to the utopia point was used as a metric to select a single design. By

plotting each electrode design according to its objective functions, the design closest to the utopia

point was selected for analysis. The utopia point is an idealized location where all objective

functions are simultaneously minimized. In this case, the utopia point was defined as the location

Page 99: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

70

where all the objective functions simultaneously equal zero. As a metric to the compare optimal

solutions from Case 1 and Case 2. The distance to the utopia point (𝑑𝑢) was defined as shown in

Eqn. 2.10.

𝑑𝑢 = √𝑓𝑜𝑣𝑒𝑟2 + 𝑓𝑢𝑛𝑑𝑒𝑟

2 (2.10)

This metric is unitless and represents a measure of how closely the ablation zone matches

the target shape.

2.3.3 Results from the Systematic Optimization Approach

For the purpose of the EUS-RFA electrode for pancreatic cancer, the materials used in

the TAM were assumed to be pancreatic tissue, polyurethane for the insulated shaft, and nickel-

titanium (Nitinol) for the exposed electrode base, outer tines, and center tine. Nitinol was chosen

for its superelasticity and biocompatibility. For Nitinol, the superelastic property means there is

minimal increase in stress for a large increase in strain during the phase transition from austenite

to martensite. The material properties used for each of the above listed materials are specified in

Table 2.2 [347–352]. The target ablation zone was assumed to be a 25 mm sphere, based on

typical tumors that would be treated with such a procedure. Each simulation was for an ablation

procedure time of six minutes. A voltage 𝑉0 = 20 Volts was applied to the electrode. An ambient

temperature 𝑇∞ = 37°C was chosen based on standard body temperature. The size of the

pancreatic tissue used during the simulation was a cylinder of radius 𝑟𝑡 = 6cm with a height ℎ𝑡 =

12cm, for which boundary effects from the proximity of grounding pads are negligible. The

convergence of the TAM was performed for a set of electrode design parameters by continuously

refining the mesh until the ablation zone volume remained essentially unchanged (within 0.5%).

The insulated shaft was represented as a cylinder of radius 𝑟𝑏 = 0.48 mm with a height ℎ𝑏 = 5 cm,

Page 100: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

71

centering the electrode in the pancreatic tissue. The length of the exposed electrode base which

connects the outer tines to the central tine was held fixed at 1 mm.

Table 2.2 Material properties for TAM

Material

Density

[𝑘𝑔 𝑚3⁄ ] Thermal

Conductivity [𝑊 (𝑚 ∙ 𝐾)⁄ ]

Electrical

Conductivity [𝑆/𝑚]

Heat

Capacity [𝐽 (𝑘𝑔 ∙ 𝐾)⁄ ]

Ambient

Temperature [𝐾]

Perfusion

Coefficient [1 𝑠⁄ ]

Pancreatic

Tissue[347,348] 1087 0.51 0.566 3164 37 --

Blood[347,349] 1050 -- -- 3617 37 0.0064

Polyurethane[350,35

1] 70 0.026 1x10-8 1045 -- --

Nitinol[352] 6450 18 1.25x106 836.8 -- --

Two cases are presented to illustrate how the optimization technique can be useful in

systematic design of EUS-RFA electrodes. Each case is summarized in Table 2.3. Case 1, shown

in Figure 2.3A-C, presents an electrode design in which there were eight outer tines and each tine

has the same length and radius of curvature. In keeping each of the outer tines the same length

and curvature, the number of design variables was reduced to three: 𝑙𝑐, 𝑙𝑜, and 𝑟𝑜. Case 2, shown

in Figure 2.3D-F, allows more freedom to the optimization than Case 1 by considering more

complex electrode designs. The number of tines was increased to 12 and there were two

configurations of outer tines, such that every other tine had the same configuration. All odd

number tines had the same length (𝑙𝑜𝑜𝑑𝑑) and radius of curvature (𝑟𝑜

𝑜𝑑𝑑) while all even number

tines had the same length (𝑙𝑜𝑒𝑣𝑒𝑛) and radius of curvature (𝑟𝑜

𝑒𝑣𝑒𝑛). The upper and lower bounds for

each design variable are shown in Table 2.3. They were chosen such that a wide range of

electrode designs could be simulated, from a long and thin profile to a very short and round

profile.

Page 101: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

72

Table 2.3 Case 1 and Case 2 optimization and electrode design parameters

Case 1 Case 2

Design parameters 3 5

Stowed diameter (𝑑𝑒) 0.96 mm 0.96 mm

Center length (𝑙𝑐) 𝑙𝑐 𝑙𝑐 Number of tines (𝑛𝑡) 8 12

Outer tine length (𝑙𝑜𝑖 ) 𝑙𝑜

𝑖 = 𝑙𝑜 , 𝑖 = 1, … ,8 𝑙𝑜𝑖 = {

𝑙𝑜𝑜𝑑𝑑 , 𝑖 = 1, 3, … , 11

𝑙𝑜𝑒𝑣𝑒𝑛 , 𝑖 = 2, 4, … ,12

}

Outer tine radius (𝑟𝑜𝑖) 𝑟𝑜

𝑖 = 𝑟𝑜 , 𝑖 = 1, … ,8 𝑟𝑜𝑖 = {

𝑟𝑜𝑜𝑑𝑑 , 𝑖 = 1, 3, … , 11

𝑟𝑜𝑒𝑣𝑒𝑛 , 𝑖 = 2, 4, … ,12

}

Exposed electrode base (𝑙𝑏) 1.0 mm 1.0 mm

Constraint parameter (𝐶𝑙𝑖) 0.8 0.8

Stall generations (𝑛𝑠𝑡𝑎𝑙𝑙) 5 5

Pareto change tolerance (𝜂) 0.001 0.001

Grid Points (𝑛𝑔𝑝) 23 23

Target radial distance (𝑡𝑖) 12.5 mm 12.5 mm

Optimization Bounds and

Constraints

{5 𝑚𝑚5 𝑚𝑚5 𝑚𝑚

} ≤ {𝑙𝑐𝑙𝑜𝑟𝑜

} ≤ {25 𝑚𝑚25 𝑚𝑚200 𝑚𝑚

}

−𝑙𝑐 + 0.8𝑙0𝑖 ≤ 0

{

5 𝑚𝑚5 𝑚𝑚5 𝑚𝑚5 𝑚𝑚5 𝑚𝑚}

{

𝑙𝑐𝑙𝑜𝑜𝑑𝑑

𝑙𝑜𝑒𝑣𝑒𝑛

𝑟𝑜𝑜𝑑𝑑

𝑟𝑜𝑒𝑣𝑒𝑛}

{

30 𝑚𝑚30 𝑚𝑚30 𝑚𝑚200 𝑚𝑚200 𝑚𝑚}

−𝑙𝑐 + 0.8𝑙0𝑒𝑣𝑒𝑛 ≤ 0

−𝑙𝑐 + 0.8𝑙0𝑜𝑑𝑑 ≤ 0

The objective functions 𝑓𝑜𝑣𝑒𝑟 and 𝑓𝑢𝑛𝑑𝑒𝑟 were based on grid points around the ablation

zone surface; see Figure 2.5. These points correspond to positions at θ = 0°, 45°, and 90°. At each

θ location (0°, 45°, and 90°), 8 positions were calculated for corresponding to φ = 0°, 22.5°, 45°,

67.5°, 90°, 112.5°, 135°, 157.5°, and 180°. Repeated grid points at φ = 0° and 180°, for θ = 0°,

45°, and 90°, were removed, resulting in a total of 23 grid points to determine the objective

functions. The target radial distance (𝑡𝑖) was equivalent for each point around the surface (𝑡𝑖 =

12.5 mm). Convergence was determined by an average relative spread change of less than 0.1%

over the final 5 generations (𝑛𝑠𝑡𝑎𝑙𝑙 = 5, 𝜂 = .001).

Page 102: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

73

2.3.3.1 Case 1

The optimization was performed using a population of 75 designs per generation and

converged after 138 generations. Figure 2.6 shows the entire performance space, i.e., a

comparison of the objective function values for each design evaluated. Each point on the plot

represents a feasible design, the plot shows the tradeoff between objective functions which exist

for some designs. From the objective functions, one can quickly understand the relative size and

shape of the ablation zone. For example, for designs along the horizontal axis (𝑓𝑜𝑣𝑒𝑟 = 0), all 23

grid points were within the treatment zone. Other designs along the Pareto front decrease 𝑓𝑢𝑛𝑑𝑒𝑟,

or fill more of the target volume, but also treat some regions of tissue beyond the target zone.

This range of optimal designs allows for discretion of whether it is preferable to only treat tissue

within the target region or to increase volume of treated tissue within the target region at the

expense of destroying other tissue outside the target region. In this study, a single optimal

electrode design was selected for further consideration based on the design located closest to the

utopia point.

Page 103: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

74

Figure 2.6 Each design tested during the optimization is plotted in this figure according to its

corresponding objective function values, adapted from [339]

For Case 1, the parameters for the optimal electrode design selected for comparison

against Case 2 were 𝑙𝑐, 𝑙𝑜, and 𝑟𝑜 equal to 18.33 mm, 15.57 mm, and 10.42 mm, respectively,

with 𝑓𝑜𝑣𝑒𝑟 and 𝑓𝑢𝑛𝑑𝑒𝑟 equal to 0.0 and 0.1027, respectively. The distance from the utopia point to

the optimal design was 0.1027. The temperature profile and ablation zone for the optimal design

from Case 1 are shown in Figure 2.7. The electrode design parameters for Case 1 are summarized

in Table 2.4 for comparison against Case 2.

Page 104: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

75

Figure 2.7 (Left) The side view of the temperature profile for an optimal electrode solution for Case

1. The ablation zone is marked by the white isothermal contour (60°C). (Right) The ablation zone

surface (green) is shown compared to the target shape (black outline) for the top and side views,

adapted from [339].

2.3.3.2 Case 2

The optimization was performed using a population of 75 designs per generation and

converged after 91 generations. The entire design space is shown in Figure 2.8. As with Case 1, a

wide range of optimal designs lie along the Pareto front providing a variety of options for optimal

designs. From the final Pareto front of designs, the parameters for the optimal electrode design

selected for comparison against Case 1 were 𝑙𝑐, 𝑙𝑜𝑜𝑑𝑑, 𝑟𝑜

𝑜𝑑𝑑, 𝑙𝑜𝑒𝑣𝑒𝑛, and 𝑟𝑜

𝑒𝑣𝑒𝑛 equal to 22.42 mm,

19.49 mm, 20.31 mm, 15.39 mm, and 8.49 mm, respectively, with 𝑓𝑜𝑣𝑒𝑟 and 𝑓𝑢𝑛𝑑𝑒𝑟 equal to

0.0434 and 0.0555, respectively. The distance from the utopia point to the optimal design was

0.0705. The temperature profile and ablation zone for the optimal design from Case 2 are shown

(Axis units are shown in meters)

Side View

Top View

Page 105: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

76

in Figure 2.9. The electrode design parameters for Case 2 are summarized in Table 2.4 for

comparison against Case 1.

Figure 2.8 Each Case 2 design simulated is plotted above according to its corresponding objective

function values, adapted from [339]

Page 106: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

77

Figure 2.9 (Left) The side view of the temperature profile for an optimal electrode solution for Case

2. The ablation zone is marked by the white isothermal contour (60°C). (Right) The ablation zone

surface (green) is shown compared to the target shape (black outline) for the top and side views,

adapted from [339].

Table 2.4 Electrode designs selected for comparison from the Case 1 and Case 2 Pareto-optimal

solution sets

Center Length Outer Length Outer Radius

Case 1 𝑙𝑐(mm) 𝑙𝑜(mm) 𝑟𝑜(mm)

Optimal 18.33 15.57 10.42

Case 2 𝑙𝑐(mm) 𝑙𝑜𝑜𝑑𝑑(mm) 𝑙𝑜

𝑒𝑣𝑒𝑛(mm) 𝑟𝑜𝑜𝑑𝑑(mm) 𝑟𝑜

𝑒𝑣𝑒𝑛(mm)

Optimal 22.42 19.49 15.39 20.31 8.49

Top View

Side View

Page 107: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

78

2.3.4 Discussion and Comparison to Standard Straight Electrode

A systematic design optimization approach was useful in comparing electrode designs by

reducing the amount of trial and error and intuition-based experimental work. A comparison of

the two electrode designs in Cases 1 and 2 can be used to show the benefit of such a systematic

approach.

The purpose of Cases 1 and 2 was to compare two electrode designs, where the second

was a more complex design, which increases freedom of the optimization to find an electrode

which generates an ablation zone that matches the target, a 25 mm sphere. For Case 1, all the

outer tines have the same length and radius of curvature, limiting spherical ablation zones which

can be generated. In particular, as the size of the target shape increases, it becomes difficult for

the electrode to fill the target shape while all outer tines have the same configuration. This can be

seen in Figure 2.7, where the ablation zone could be improved by broadening the lower and upper

portions. With a single outer tine configuration, the electrode was unable to fill these lower and

upper portions. Increasing the number of tines and allowing multiple configurations should allow

for better shape matching than only eight outer tines with a single configuration.

For Case 2, Figure 2.9 shows that increasing the number of outer tines and allowing for

two outer tine configurations resulted in significant improvement in the ability of the electrode to

cover the target zone. The optimization improved the ablation zone by using the second set of tine

configurations. As evident from the objective function values, the optimization was able to find a

design with a decreased distance to the utopia point in Case 2. Thus, the optimization approach

was able to show that the Case 2 design has the potential for improved shape matching compared

to Case 1. Depending on the increase in complexity for manufacturing, the improved objective

function values of Case 2 may be desirable. For complex, non-symmetrical, target shapes, it is

expected that deployable electrode designs that allow even more freedom than Case 1 or Case 2

Page 108: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

79

will be necessary. The objective functions utilized for this optimization were useful for matching

a sphere and can easily be expanded to complex or arbitrary shapes.

To compare against currently available endoscopic electrodes, the ablation zone for a

standard straight needle is shown in comparison to the Case 1 and Case 2 solutions; see Figure

2.10. The length of the straight electrode is typically adjustable up to 20 mm. In this case, the full

electrode length (20 mm) was used for simulation and comparison. By comparing the ablation

zone volumes with the target shape, one can obtain a quantitative sense of the improvements seen

with the optimal designs. The percentage of the target volume filled by the ablation zone was

computed by dividing the ablation zone volume inside the target zone by the total target volume.

The percentage of the ablation zone outside the target volume was computed by dividing the

ablation zone outside the target volume by the total ablation zone volume. The current straight

electrode ablated only 25.1% of the target volume. The optimal solutions show increased

treatment volumes of 70.9% and 87.0% for Case 1 and Case 2, respectively. The results for the

straight electrode and optimal solutions for Case 1 and Case 2 are summarized in Table 2.5.

Table 2.5 Quantitative comparison of a straight electrode and optimal Case 1 and Case 2 solutions

Single Case 1 Case 2

Percentage of target zone filled 25.1% 70.9% 87.0%

Percentage of ablation volume outside target zone 0.5% 0.4% 3.2%

Page 109: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

80

Figure 2.10 Top and side views of the ablation zones for the straight electrode and optimal solutions

of Case 1 and Case 2. The green surface represents the 60°C isothermal surface and the black circle

represents the target zone, adapted from [339].

While this optimization approach has shown promise as a useful tool in simulating the

ablation zone and designing RFA electrodes, limitations exist in accounting for the numerous

variables of the human body. Experimental validation of the TAM using a thermochromic tissue

phantom is described in Section 2.4.3. When compared with other similar models reported in

literature, the average error between temperature profiles across the electrode was found to be

1.8% [61]. In addition, the pancreatic tissue in this TAM does not account for the inhomogeneity

which exists in all biological tissue; however, the random and constant dividing of cancer cells

produces more homogeneous tissue as compared to typical healthy tissue. The tissue properties

Page 110: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

81

were also chosen based on an approximation of tumors, though, in reality, a variety of values are

reported in literature for various pancreatic tissue properties.

One challenge with selecting tissue properties was that the reported values in literature

vary and are often temperature or frequency dependent. For this optimization, the electrical

conductivity and blood perfusion were assumed to be constant according to reported values for

pancreatic tissue. The electrical conductivity is temperature dependent as discussed by Chang

[71,347] and has been explored in many other studies [63,348,353]. The blood perfusion rate has

also been shown to decrease with increased metabolic activity in tumors making it difficult to

determine an appropriate value [349]. Although the overall size and temperature of the ablation

zone was found to be sensitive to these parameters, the shape of ablation zone was not heavily

affected. Thus, the current TAM could be used to understand the general shape of the ablation

zone and then tissue properties can be used to tune the TAM to match experimental ex-vivo

testing for validation. It is also important to recognize that RFA may have limited effectiveness

near large blood vessels because they act as a significant heat sink, limiting the temperature to

which the tissue can be heated. While this TAM does not include the heterogeneities present in

living tissue (blood vessels, surrounding organs, or other structures), the same optimization

approach could be utilized with models of heterogeneous tissue to generate an electrode design

for patient specific treatment planning.

While this approach for design optimization of RFA ablation electrodes shows great

promise, the TAM only accounts for an electrode shape after it has been inserted in the tissue.

The deployment of the electrode, and therefore the ablation zone produced, were likely to be

affected by variations in tissue elastic modulus and other tissue properties. To consider the

robustness of the current electrode designs, the outer tine radius of curvature was varied by ±10%

for the Case 1 solution and the ablation zone volume was calculated. For an outer tine radius of

curvature of 9.38 mm, the ablation zone volume increased by 3.3%. For an outer tine radius of

Page 111: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

82

curvature of 11.46 mm, the ablation zone volume decreased by 3.7%. Therefore, for slight

variations in outer tine radius of curvature, the change in total ablation zone volume was small.

To account for these variations in deployment, it is necessary to develop modeling which predicts

the deformation and mechanical response of the electrode as it deploys while being inserted into

the tumor. This type of deployment model will be useful in relating the manufactured electrode

shape with the final electrode shape deployed in tumors.

2.3.5 Conclusions from Systematic Optimization Approach

The purpose of this study was to optimize the shape of an RFA electrode to match the

ablation zone for a target tumor geometry to improve the efficacy of treatment of pancreatic

cancer. Using a computational model based on coupling the electric and bioheat transfer

equations in COMSOL Multiphysics software, a TAM was used to predict the ablation zone for

unique electrodes. The resulting electrode shape was optimized to generate a 2.5cm spherical

ablation zone using a multi-objective genetic algorithm. For a target 2.5cm spherical tumor, the

optimal design parameters of the compliant electrode design were found for two cases. Case 1

and Case 2 optimal solutions filled 70.9% and 87.0% of the target volume as compared to only

25.1% for a standard straight electrode. While the TAM and electrode were specifically focused

on EUS-RFA of pancreatic cancer, RFA is a common procedure for tumors throughout the

abdomen. This approach of optimizing the electrode shape for a target ablation zone has the

potential to personalize treatment for many cancer patients and may be applied to a variety of

tissues where RFA is practical.

Page 112: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

83

2.4 Testing and Validation of Deployable Electrode

2.4.1 Prototype Development

An early challenge in the development of the deployable EUS-RFA electrode was size

restrictions due to the requirement that the deployable electrode pass through an endoscopic

needle (~1 mm diameter). The number of manufacturing processes that can produce such a small-

scale device was limiting but ultimately two different processes were used for early prototypes:

metal AM with laser powder bed fusion (L-PBF) and laser micromachining (LMM).

L-PBF was used to create a 2x scale version of the electrode, printed on the 3D Systems

ProX 320 using Oerlikon MetcoADD 625A powder at Penn State’s CIMP3D facility. The parts

were removed from the build plate using wire electrical discharge machining and post treated

with hot isostatic pressing to reduce voids and stress relieve the part. Stowing the L-PBF

prototype inside the sheath, inner and outer diameter 2.16 and 2.41 mm, respectively, resulted in

some visible plastic deformation. Next the base of the part was welded onto a steel wire with a

diameter of 0.991 mm. The final part dimensions of the L-PBF prototype are shown in Figure

2.11. These prototypes were used in the deployment testing in a tissue phantom described in

Section 2.4.2. An additional version of the L-PBF prototype with only six outer tines was created

for testing the limits of the manufacturing process and resulted in a nearly 1x version; see Figure

2.12.

Page 113: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

84

Figure 2.11 L-PBF 2X scale electrode prototype

Figure 2.12 Near 1X scale L-PBF electrode prototype

The second manufacturing method, LMM, used a superelastic Nitinol tube to create the

outer compliant tines. Initial prototypes were manufactured by L4is. The Nitinol tube had an

outer diameter of 0.991 mm and an inner diameter of 0.584 mm. A nanosecond laser was used to

make 14.8 mm cuts along the length of the tube starting at the end of the tube. In total four cuts of

this length were equally spaced around the tube circumferentially creating 4 tines. The final width

of each tine was 0.635 mm. The tip of each tine was also sharpened with two cuts from the

nanosecond laser to make the tip of each tine pointed. Finally, each electrode tine was wrapped

49.8 mm

15.4 mm

Page 114: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

85

around a mandrel to form it into its naturally deployed configuration. These outer compliant tines

were fitted onto a steel wire with a diameter of 0.584 mm which serves as the stiff central needle.

The final dimensions of the electrode fabricated with LMM are shown in Figure 2.13. These

prototypes were used in the deployment testing in a tissue phantom described in section 2.4.2.

Since that time, additional prototypes were fabricated based on the Case 1 and Case 2 optimal

configurations described previously, featuring 8 and 12 outer tines, respectively; see Figure 2.14.

These prototypes were fabricated through collaboration with Actuated Medical Inc. (AMI) with

LMM by Spectralytics from a Nitinol tube with an outer diameter of 0.030” ± 0.001” and an inner

diameter of 0.019” ± 0.001”. These later prototypes were used in the ex-vivo experimentation for

validation of the TAM.

Figure 2.13 4-tine laser micromachined prototype

19.6 mm

13.2

mm

Page 115: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

86

Figure 2.14 8-tine and 12-tine prototypes based on optimal solution configurations that used for ex-

vivo experimentation

2.4.2 Deployment in a Tissue Phantom

The purpose of the deployment testing was to provide proof of concept for the deployable

EUS-RFA electrode and to determine how insertion speed affects the insertion force and amount

of deployment in the two electrodes when inserted into a tissue phantom. The deployment testing

was completed with the help of undergraduate students Katherine Reichert and Fariha Azhar who

assisted with the experimental setup and testing.

2.4.2.1 Methods

The test setup, shown in Figure 2.15, includes an Aerotech ECO115SL linear slide,

controlled by a Soloist MP10 controller. A Futek LSB302 single axis load cell was used to

measure the amount of force required during insertion. The electrode was positioned in line with

the load cell that was fixed to the slide. LabView software was used to control the insertion speed

and distances while acquiring data from the load cell.

Page 116: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

87

Figure 2.15 Insertion force and deployment experimental setup for early prototypes of the EUS-RFA

electrode

A Polyvinyl Chloride (PVC)-based tissue phantom was prepared using M-F

Manufacturing Company’s regular liquid plastic and softener, mixed with a ratio of 4:1. This ratio

was chosen from literature on flexible needle insertions which reported the phantom elastic

modulus similar to porcine tissue [340]. The tissue phantom was placed on an elevated platform

to align it with the load cell and electrode. The tip of the sheath for the electrode was partially

inserted into the phantom and fixed to the platform. The electrode prototype was initially

compressed and stowed inside the distal end of the sheath and connected in line with the load cell.

As the slide was advanced, the compressed electrode slid through the sheath and deployed into

the tissue phantom. After each trial, the position of the tissue phantom was adjusted on the

platform to prevent overlap in the deployment region with previous trials.

The force and deployment of LMM and L-PBF electrodes were measured at speeds of 1,

3, 5, 7, 10, 25, 35, and 50 mm/s. To minimize noise in the force data, a Savitzky-Golay

Page 117: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

88

smoothing filter was applied to the data in MATLAB. The filter successively fits polynomials to

the data using a method of linear least squares. The force for the L-PBF electrode was measured

along 80 mm of travel which was split into three segments. Segment 1 was 36 mm and represents

the friction force associated with the electrode sliding through the sheath. It ends when the central

needle first enters the tissue phantom. Next, segment 2 was 10 mm and ends when the outer tines

leave the sheath and pierce the tissue phantom. Finally, segment 3 was 34 mm and ends when the

electrode was fully inserted. The force for the LMM electrode was measured along 40 mm of

travel which was be split into three segments using a method similar to the L-PBF electrode.

Segment 1 was 20 mm, segment 2 was 4 mm and finally, segment 3 was 16 mm.

To determine the amount of deployment, a digital photo was taken after each trial to

record the spread, or maximum width of the electrode tines. The photos were analyzed using

digital image correlation to determine the average spread at different speeds.

2.4.2.2 Results

The force, position and amount of deployment readings collected for the L-PBF and

LMM electrodes were plotted and analyzed using MATLAB and ImageJ [354] image analysis

software.

The average spread for the L-PBF prototype at the speeds of insertion is shown in Figure

2.16 (left). Error bars denote one standard deviation from the average. The average spread at

insertion speeds of 1, 3, 5, 7, 10, 25, 35, and 50 mm/s were 12.3, 14.5, 12.1, 12.3, 14.8, 12.5,

12.4, 8.6 mm, respectively. The testing shows a maximum spread around 10 mm/s. The insertion

force as a function of the slide position for each speed was then averaged over the five trials for

each of the 8 speeds. Figure 2.16 (right) shows a comparison of the average insertion force profile

Page 118: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

89

for each of the insertion speeds for the L-PBF electrode. Here, a zero position corresponds to the

point at which the tip of the electrode first penetrates the tissue phantom.

Figure 2.16 (Left) The spread or amount of deployment is shown based on the insertion speed.

(Right) The insertion force profile is separated into three segments showing the friction sliding

through the sheath, the penetration of the central needle, and then the deployment of the outer tines.

Both graphs are for the L-PBF electrode.

Similarly, the average spread for the LMM prototype was plotted against speeds of

insertion and is shown in Figure 2.17 (left). Error bars denote one standard deviation from the

average. The average spread at insertion speeds of 1, 3, 5, 7, 10, 25, 35, and 50 mm/s were 16.7,

19.2, 17.5, 15.5, 19.9, 17.5, 14.9, 14.8 mm/s, respectively. The testing shows a maximum spread

to be around 10 mm/s which was the same as the L-PBF electrode. The insertion force as a

function of the slide position for each speed was averaged over the five trials. Figure 2.17 (right)

shows a comparison of the average insertion force profile for each of the insertion speeds. Here a

zero position corresponds to the point at which the tip of the electrode first penetrates the tissue

phantom.

Page 119: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

90

Figure 2.17 (Left) The spread or amount of deployment is shown based on the insertion speed.

(Right) The insertion force profile is separated into three segments showing the friction sliding

through the sheath, the penetration of the central needle, and then the deployment of the outer tines.

Both graphs are for the LMM electrode.

2.4.2.3 Discussion

The data collected reveals some interesting trends for future development of deployable

endoscopic electrodes. In addition, important limitations and advantages due to the manufacturing

process became apparent.

Spread was at a maximum around 10 mm/s and appeared to decrease as insertion speed

increases beyond 10 mm/s; see Figure 2.16 (left). The decrease in average spread for the L-PBF

electrode from 10 mm/s to 50 mm/s was nearly 60%. However, in the case of the LMM electrode,

a decrease in spread can be observed from 10 mm/s to 50 mm/s, though the decrease was less

than for the L-PBF electrode. This could be because of the considerable difference in size

between the two prototypes. At speeds lower than 10 mm/s the data did not show a clear trend for

either electrode. For some speeds the error bars were large and were attributed to the complexities

of surface friction and the viscoelasticity of the tissue phantom. There appears to be an ideal

Page 120: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

91

speed to maximize spread. It was likely that the viscoelasticity and the speed dependent friction

behave differently across the range of speeds tested. For example, at slow speeds the tissue

phantom was able to spring back quickly relative to the insertion speed which affects the overall

deployment.

During the insertion, the tissue phantom deforms as the central needle tip pushes on it.

After a certain point, the tip pierces and ruptures the tissue phantom. The rupture allows the tissue

phantom to spring back toward its original position. At slow speeds, the insertion force remains

low because the tissue phantom was allowed to spring back, reducing the potential energy like a

spring. This spring back behavior was visible twice along insertion force profiles. The first spring

back behavior occurred after deformation from the central needle and the second occurred after

deformation from the outer tines. The spring back is most clearly visible in segment 3 of Figure

2.16 (right) (L-PBF) and Figure 2.17 (right) (LMM), for the 1, 3, 5 and 7 mm/s force profiles. At

high speeds, the tissue phantom was not able to spring back as much before it was pushed further

and further, preventing the force from dropping as much. By segmenting the force profiles, it was

clear that the slope of the force increased as the electrode was inserted. Yet, while the slope of the

force profiles was similar across different speeds, the spring back in the tissue at slow speeds

results in low overall forces. It was interesting to note that the spring back behavior was very

pronounced in the LMM electrode, with the force dropping almost to the force before the initial

penetration. This was likely due to the size and surface roughness of the LMM electrode

compared to the L-PBF electrode. Because the L-PBF electrode had a rougher surface finish and

was larger than the LMM electrode, the friction forces were higher and did not allow the tissue to

spring back as much as for the LMM electrode.

The L-PBF manufacturing processes presents a unique opportunity for fabrication of

electrodes with complex geometry. While the testing results reported were for a 2x scale

prototype, recent work has shown that the electrode may be printed near actual size; see Figure

Page 121: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

92

2.12. The primary limitation with the L-PBF process was the material selection. Inconel is not

biocompatible due to the high nickel content and lacks the superelastic property exhibited by

Nitinol. In addition, the initial L-PBF prototype after the HIP post processing exhibited spread of

21.4 mm, requiring a significant amount of plastic deformation to be stowed in the sheath. After

being stowed in the sheath, the spread in air without resistance was decreased to 15.4 mm. Plastic

deformation negatively impacts the deployment potential by limiting the spread in air and the

amount of elastic strain energy that can be stored when compressed into the sheath. While other

biocompatible material choices exist for L-PBF, the ideal material with a superelastic behavior to

resist plastic deformation, Nitinol, is in the early stages of process parameter development on

metal printers.

For the LMM electrode, there were some clear advantages in the inherent surface

roughness as compared to the L-PBF electrode, though the forming process to bend the outer

tines following the micromachining process was challenging because of the large strains that the

Nitinol can undergo prior to plastic deformation. A special forming technique with a mold or

another method based on a mandrel would have to be devised to improve the consistency.

Additionally, the shapes of the electrodes would be difficult to customize as compared to the L-

PBF electrode in which the complexity was not a major issue.

2.4.2.4 Conclusions

While some trends can be seen in the data presented, there were several limitations that

should be taken into account. First, the tissue phantom was chosen based on matching elastic

modulus properties for healthy porcine tissue, but other properties such as friction could play a

role in deployment. Second, the tissue phantom was homogeneous. In reality, a tumor consists of

a heterogeneous fibrotic structure. Third, the L-PBF prototype tested was a 2x scale version. With

Page 122: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

93

these limitations in mind, the authors expect the trends in the spread and insertion force to remain

similar across various elastic modulus values for tissue, heterogeneous compositions, and scales.

In conclusion, it has been shown that the speed of insertion results in minor variations in

the amount of deployment. The insertion speed should be controlled for consistency during use.

Furthermore, while AM was able to produce functional metal parts with complex geometries,

there are limited materials that can be printed at such scales that allow for the compliance needed

to stow and deploy effectively.

2.4.3 Ex-vivo validation of TAM

The purpose of the ex-vivo experimentation was to validate the TAM in predicting the

ablation zone surrounding the electrode. The experimental validation of the TAM was performed

using the 8-tine prototype shown in Figure 2.18; see also Figure 2.14. The prototype was

fabricated using laser micromachining of a Nitinol tube. Next, a machined tube was fitted onto a

beveled stainless-steel wire that serves as the stiff central needle. The shape of the deployed

electrode was defined by the length of the center needle, length/curvature of the outer tines, and

length of the conductive base; see Figure 2.18.

Page 123: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

94

Figure 2.18 Geometry details for 8-tine deployable electrode prototype for TAM validation, adapted

from [355]

For validation of the TAM, the thermal and electrical material properties of the

cylindrical tissue block were selected to match the material properties of the thermochromic

tissue phantom used in the experimental setup; see Section 2.4.3.1. For the TAM, the electrode

was modeled in its deployed state and its conductive surfaces were applied a constant applied

voltage. The applied voltage levels were 26.4V and 17.5V, corresponding to the high and low

power levels in the tissue phantom model, respectively. The voltage levels were calculated using

a known power level and measured resistance. As described in more detail in Section 2.3.1, the

60°C temperature contour was used to mark the ablation zone for the TAM. The size and shape of

the 60°C contour predicted by the TAM was validated by comparison to the thermochromic

tissue phantom.

Ex-vivo experimentation was performed with assistance from our collaborators at AMI

and undergraduate student Fariha Azhar in development of the experimental setup and testing.

Page 124: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

95

2.4.3.1 Methods

Tissue phantom models were fabricated to visualize thermal tissue damage [356]. In this

work, a thermochromic tissue phantom for tumor ablation was chosen as the material for

experimental validation of the electrode [357]. The phantom was based on a polyacrylamide gel

that mimics tissue. The thermochromic agent was selected so that the phantom permanently

turned a deep shade of magenta as the temperature rises above 60°C, the same temperature that

causes instantaneous cell death (i.e. ablation or necrotic zone). The ingredients, percent volume

for each ingredient, and fabrication method for the phantom were consistent with Negussie et al.

[357], but the total volume was modified. The thermochromic tissue phantom was molded in

cylinders with a 3.65 cm diameter and a 7.62 cm height. Electrical conductivity of the tissue

phantom was measured, and the other material properties of the tissue phantom were from

Negussie et al. [357]; see Table 2.6.

Table 2.6: Tissue phantom material properties [357]

Material Property Value

Density 1033 kg/m3

Heat Capacity 3939 J/kg∙K

Thermal Conductivity 0.59 W/m∙K

Electrical Conductivity 0.587 S/m

The thermochromic tissue phantom experiments were carried out in a custom tank

containing warmed (~37°C) physiologic (0.9%) saline; see Figure 2.19. Prior to test, the phantom

sample was placed in the saline until warmed to within several degrees of tank temperature. The

outer electrode sheath (with un-deployed electrode inside) was inserted into the phantom to a

depth that ensured the entire electrode was contained in the phantom sample once deployed. After

deploying the electrode, the sheath was retracted to fully expose the electrode while avoiding

electrical contact with it. To monitor temperatures of the phantom during ablation, a custom 7-

Page 125: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

96

channel thermocouple needle array was inserted at the level of the electrode to a depth of ~1.5

cm, being careful not to contact the electrode tines. An 8th thermocouple was used to monitor tank

temperature. All thermocouples were connected to a multi-channel thermocouple reader (TC-08,

Omega) and output was displayed on a laptop during the experiment. A patient ground pad

(hydrogel coating removed) was submerged and secured to one end of the tank with adhesive

tape.

Figure 2.19 Schematic and photo of the experimental set-up for thermochromic tissue phantom

testing, adapted from [355]

A clinical RF ablation generator (model 950, Aaron) was connected to the electrode and

ground pad. The generator was operated in “blend” mode (357 kHz RF in short bursts delivered

at a rate of 30 kHz). Two different power levels (15 W and 20 W) were tested in ablation trials

lasting 15 minutes. A 2-channel digital oscilloscope (TDS1000, Tektronix) was used to monitor

voltage signals on both the electrode and ground pad (oscilloscope channel grounds connector

together, floating). Additionally, an LCR meter (7600 Model B, QuadTech) was used to measure

impedance (at 357 kHz) between the electrode and ground pad before/after ablation trials.

Page 126: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

97

Three repetitions of each power level (15W, 20W) in a different tissue phantom sample

were used to determine an average ablation zone for each power level. After conducting the

experiments, each tissue phantom sample was sliced in half, length-wise along the cylinder.

Preliminary studies showed the ablation zone to be axisymmetric so a single cross section through

the center of the cylinder, length-wise, was used to compare the experimental and TAM ablation

zones. The ablation zone, marked by the thermochromic color change, was imaged and analyzed

using ImageJ and compared to TAM results in MATLAB.

Digital images for each trial were scaled and calibrated based on the resolution of the

image and a known distance in the image. An ellipse was fit to each ablation zone to approximate

the size and shape of the ablation zone. The ellipse was fit by matching the normalized second

central moment to those of the ablation zone, using regionprops, a MATLAB function. Similar to

the experimental results, an image of the 60°C contour from the TAM was generated for the low

and high-power setting. The measured major and minor axes of the fit ellipse for each trial were

averaged for the respective power level and compared against the measured axes in the simulated

ablation zones from the TAM.

2.4.3.2 Results

Figure 2.20 and Figure 2.21 show examples of the original and processed images for the

experimental and TAM ablation zones at 15W and 20W, respectively. In these figures, the dark

pink and lime green are the ablation zones. Using image analysis, the white profile was the

ablation zone and the red ellipse was fit to approximate the size and shape by the major and minor

axes. The average major and minor axes of the elliptical fits to ablation zones for the tissue

phantom experiments (3 trials per power level) and the TAM are plotted in Figure 2.22. The error

bars denote the standard deviation for major and minor axes. See discussion below for possible

Page 127: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

98

reasons for disparities between the experimental and TAM-generated results. A second set of

simulation results are shown to demonstrate the effect of differing electrical conductivity of the

electrode base and tissue phantom as a way to model tissue damage; see “Adapted” in Figure

2.22. These varied conditions account for differences between the ex-vivo and TAM models and

are detailed in the discussion.

Page 128: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

99

Figure 2.20: Tissue phantom and TAM ablation zone and analyzed images for the 15W power level,

adapted from [355]. (Scale 15.4 px/mm)

Figure 2.21: Tissue phantom and TAM ablation zone and analyzed images for the 20W power level,

adapted from [355]. (Scale 17.4 px/mm)

Page 129: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

100

Figure 2.22: Comparison of the ablation zone size and shape by the major and minor axis of an

ellipse fitted to each ablation zone with image analysis; see examples in Figure 2.20 and Figure 2.21,

adapted from [355]

2.4.3.3 Discussion

The data collected reveals interesting differences between the tissue phantom and TAM

results. In particular it was clear that the TAM overpredicted the size of the ablation zone in

comparison to the tissue phantom, particularly for the 20W power level. Two primary reasons for

the difference in ablation zone size between the two models were the tissue damage and electrical

conductivity of the electrode.

First, the TAM does not account for tissue damage such as charring or sparking that

occurs at elevated temperatures and voltages, causing decreased electrical conductivity of the

tissue. Figure 2.21 clearly shows damage to the tissue phantom resulting in smaller major and

minor axes at 20W power level than predicted by the TAM. At the 15W power level, the damage

was less apparent than for 20W. In RFA, charring or damage increases the tissue resistance,

Page 130: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

101

reducing the efficiency of the treatment by limiting the power that can be delivered to the

surrounding tissue. In the TAM, not accounting for increased tissue resistance allowed for

elevated temperatures near the electrode, resulting in larger thermal conduction than the ex-vivo

experiment. One method to correct this in the TAM was to apply a temperature dependent

thermal conductivity, however, the computational cost was high to recompute the electrical

conductivity and electric field throughout the simulation. A simplified approach was to reduce the

constant electrical conductivity of the entire tissue. Reducing the electrical conductivity of the

tissue in the TAM by fifty percent for the high-power level and five percent for the low power

levels, the TAM closely matches the ex-vivo model.

A second reason for the difference in the ablation zone size was the heating near the

proximal end of the electrode. In the TAM, the initial heating in the tissue occurs near the tips of

the outer tines and central needle, and proximal end of electrode. In some of the experimental

trials, heating near the proximal electrode end was inconsistent, showing signs of reduced

electrical conductivity and increased heat sink effects. The inconsistencies were suspected to be

caused by saline solution entering the tissue phantom along the electrode path.

The difference in conductivity between the tissue phantom and TAM at the proximal end

of the electrode was best seen at the low power level where tissue damage was less significant

than the high power level; see Figure 2.20. In the TAM, heating at the proximal end of the

electrode causes a broader and longer ablation region near the bottom of the profile. In contrast,

the ablation zone in the tissue phantom comes to a narrow point. The result was that the minor

axis of the tissue phantom and TAM were similar, while the major axis was not; see Figure 2.22.

By reducing the electrical conductivity of the electrode base by 4 orders of magnitude, mimicking

the ex-vivo model, the major and minor axes become very similar between the tissue phantom

and TAM. By changing both the electrical conductivity of the tissue phantom and the

Page 131: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

102

conductivity of the electrode base, we see good agreement between the TAM and tissue phantom;

see Figure 2.22.

There were a few limitations that exist in the thermal ablation and tissue phantom

models. The TAM assumes constant material properties while properties such as electrical and

thermal conductivity were known to be temperature dependent. In addition, the TAM assumes the

initial temperature was 37°C and was constant at the exterior tissue surface. The TAM also used a

constant applied voltage, a simplification of the actual pulsed RF voltage, based on the root mean

square voltage. As mentioned previously, the TAM also did not directly account for tissue

damage or charring at elevated temperatures. Though some error between the experiment and

TAM was expected by not capturing the full physics of tissue ablation, the purpose of the TAM

was a quick approximation of the ablation zone for the formal optimization with a genetic

algorithm [339].

For the tissue phantom testing, the tank heater and controller helped to maintain a

consistent starting temperature and tissue phantom boundary temperature, though there were

small variations during testing. In addition, as described previously, in some cases the tissue

phantom did not change color in the tissue phantom experiment, which were not included in the

data analysis. This result was attributed to short circuiting when the saline solution entered the

tissue phantom and/or thermocouple array contacted the outer electrode tines during testing.

2.4.3.4 Conclusions

In conclusion, the tissue phantom experiment showed that the TAM overpredicted the

ablation zone size, likely due to not accounting for tissue damage and inaccurate prediction of

heating at proximal end of electrode. By adapting the conditions of the TAM to match the ex-vivo

experiment, the TAM was validated to approximate the ablation zone. Future work should

Page 132: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

103

consider improving the TAM to consider tissue charring or damage at elevated temperatures. By

considering a variety of power levels, a modified electrical conductivity factor for the TAM could

be determined based on the maximum temperature or applied voltage during the simulation.

2.5 Closing Remarks

Optimization of the electrode geometry with a TAM can be a valuable tool when

designing specially shaped ablation electrodes. The potential to develop unique ablation

electrodes, based on a specific target geometry, or a specific tumor, can help to customize patient

care. In this work, the target shape of the ablation zone was spherical, however, the desired

ablation zone may correlate to a patient specific tumor shape. In addition, more complex models

may be developed which incorporate patient specific blood vessel structures to optimize the

electrode while considering heat sink effects. With the ex-vivo validation completed, this method

of systematic design through optimization can be used to design specially shaped electrodes for

patient specific cases. These results provide evidence toward the benefits of combining

optimization with computational modeling and its potential impact on personalized medicine. In

the next chapter, a similar methodology of developing efficient computational models, paired

with optimization, for personalized or customized applications of AM is addressed.

Page 133: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

104

Chapter 3 Unit Cell Generation through Ground Structure Topology

Optimization

The focus of this chapter is the theory behind unit cell generation through a ground

structure topology optimization method. The chapter includes a brief summary of the relevant

background and motivation as well as descriptions of the approach. The approach includes

development of a 3D ground structure model, a formal description of the optimization, and

solution of several benchmark problems to validate the approach. The following objectives and

sub tasks are addressed in this chapter:

2. Develop three-dimensional ground structure topology optimization for unit cell DfAM

2.1. Develop 3D-parametric ground structure model

2.2. Establish systematic optimization method for unit cell generation

2.2.1. Generate set of potential objective functions and constraints

2.2.2. Quantify manufacturing considerations for metal AM

2.3. Validate optimization with benchmark problems

3.1 Background and Motivation

At the core, AM processes are based on a layer-by-layer approach for building up a part,

rather than subtractively removing material. With its layer-by-layer build process, AM offers

access to the entire build volume during fabrication. This unique advantage of AM over other

manufacturing technologies enables design complexity to be easily incorporated into components

and has led to increased interest and rapid development of AM processes, particularly in the past

decade [91,337].

Lattice structures are one example of design complexity enabled by AM, and are an

increasingly active area of research for academia and industry [160,228,250]. Whether for

lightweighting, improving energy absorption capabilities, adding functionality, or improving

Page 134: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

105

biocompatibility, lattice structures are being explored extensively for designing a component’s

mesostructure. A non-stochastic lattice structure is defined as a patterned structure based on a

repeated unit cell topology. Though the freedom of AM allows for complex and intricate lattice

structures to be manufactured, the challenge remains to be able to effectively design components

with such structures in mind.

One challenge with lattice structures is that their behavior is not well understood and are

often used without specific consideration for why a particular unit cell topology was chosen over

another. The addition of lattice structures complicates traditional designs because the unit cell

topology changes the effective bulk mechanical properties from the material properties of the

base material. For example, the assumption of isotropic behavior common to many metals is no

longer a valid assumption when considering the macroscale properties of a metal lattice structure.

Commercial software for AM makes it easy to lightweight a part by simply adding a lattice, but

does not provide any guidance for which lattice will be most beneficial for the component.

Similar to the need for topology optimization methods for overall geometric design of a

component, systematic design methods are needed for the appropriate design of lattice structures.

Topology optimization is the process of determining material placement within a design space to

achieve a specific goal or objective [282]. Chen et al. [133] describe the use of continuum-based

topology optimization to design novel isotropic lattice structures that exhibit high stiffness. Watts

et al. [313] and Watts and Tortorelli [314] have published methods for reducing computational

cost of lattice structure design through surrogate models and geometric projection methods.

While there are numerous research groups exploring methods for metamaterial design, or the

design of materials with properties not found in traditional materials (i.e. multi-functional,

negative Poisson’s ratio, self-healing), the direct incorporation of AM constraints into the

optimization has not been addressed. Liu et al. [295] review of the current state of topology

optimization and AM and they describe a need for further incorporation of design for AM

Page 135: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

106

(DfAM) limitations in topology design. In a review of structural optimization for AM, Plocher

and Panesar [336] cited integration of DfAM as a major bottleneck for a streamlined design

process.

DfAM constraints include overhangs and support structures, powder removal, and

minimum feature size. In AM the build angle is defined as the angle between the build plate and

the surface of the component being built. If this angle is less than 35 degrees, additional structures

are generally required to be built underneath the part to support it by anchoring it to the build

plate, otherwise, significant warping occurs. These support structures are frequently used in AM

components but become nearly impossible to remove for intricate and delicate lattice structures.

Overhangs are defined as a region where part extends out horizontally. If the horizontal region

reconnects after a short distance (~5 mm) it can form a bridge anchoring the material. Otherwise,

overhanging regions typically cause failures do to warping and overheating. For powder-based

AM processes, anywhere that material is not solidified, there is excess powder to be removed. For

lattice structures, one common challenge is avoiding trapping powder and the ability to remove

all the powder from the complex maze of channels that form in the lattice. The minimum feature

size is the smallest features that can be built by the AM process. For L-PBF, this is typically

based on the spot size of the laser. As lattice structures are typically very intricate, they quickly

push the limits of the minimum feature size or resolution of the AM machine. Without

incorporation of DfAM considerations into topology optimization methods, the optimized

structures may not be manufacturable or lose much of the enhancement gained from the

optimization after the component is built.

One method of topology optimization is ground structure topology optimization (GSTO),

a discretized approach for determining for an optimized topology. GSTO discretizes the design

space with a grid of nodes and connects the nodes using truss, beam, or frame elements. By

applying periodic boundary conditions to the ground structure model and considering

Page 136: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

107

homogenization techniques, it is proposed that GSTO can be used to design unit cell topologies

for a lattice structure with tailored properties. As compared to continuum-based topology

optimization, GSTO generally offers a reduced number of variables for lower computational cost.

It is particularly well-suited to AM unit cell topology design because many lattice structures are

based on an arrangement of trusses and nodes, and DfAM considerations can be directly

incorporated. While this approach has been used in the past for mesoscale design, there is a need

to expand such an approach to include additional objective functions and DfAM considerations.

With the need for systematic approaches for the design of unit cell topologies, this

chapter describes the additive lattice topology optimization (ALTO) approach. ALTO is a 3D

GSTO method for generation of novel unit cell topologies tailored for their application. The

GSTO serves as a systematic optimization method for generating novel lattice structures, offering

a fast simulation with fewer design variables than continuum-based approaches. ALTO includes

direct incorporation of various DfAM considerations such as overhangs and support structures,

powder removal, and minimum feature size. Through a variety of optimization objective,

constraint, and penalty functions, ALTO offers a systematic approach to unit cell topology design

that can be applied to a wide range of applications.

3.2 Additive Lattice Topology Optimization (ALTO)

The generation of novel unit cell topologies with DfAM and optimization were

accomplished through ALTO. First a ground structure model was applied to represent the unit

cell topology. Next, the unit cell topology was tailored for the application with multi-objective

optimization. After optimization, the solution was interpreted and verified for manufacturability.

Next the solution may be recreated in design software and ultimately fabricated with AM. An

Page 137: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

108

overview of the process is shown in Figure 3.1. MATLAB [344] code for the 3D version of

ALTO is provided in Appendix B.

Figure 3.1 Overview of Additive Lattice Topology Optimization (ALTO) to generate optimized unit

cell topologies for lattices fabricated with AM

3.2.1 Ground Structure Model

GSTO methods discretize a design space into a set of nodes and then connects nodes with

truss, beam, or frame elements. In a full ground structure, each node was connected to every other

node by an element, forming a highly redundant system. The ground structure is evaluated using

Page 138: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

109

standard finite element analysis. For optimization, the cross-sectional areas of each element are

the design variables, which are appropriately resized to minimize the objective function(s).

The ground structure approach for topology optimization is an alternative to the so-called

continuum-based topology optimization where the design space is discretized into finite elements

and the relative element densities are the design variables. The earliest references to GSTO dates

back to the 1960’s with work by Dorn et al. [286]. The methodology and formulation have been

described by Bendsoe and Sigmund [281] and Ohsaki [287] in their textbooks on topology

optimization. GSTO has since advanced to tackle a number of different challenges including

multi-material design [288], reliability [289], arbitrary 2D and 3D domains [290,291], adding and

removing nodes or elements [292], buckling and overlap [293], sizing non-patterned lattice

structures [159,224–227], and other AM considerations [294–296].

In particular, GSTO has been researched previously by Sigmund [297,298] for

metamaterial design of structures with homogenized material properties. Since the work by

Sigmund in the late 1990’s, others have produced similar work, generally using the approach to

develop novel materials matching a specific constitutive matrix. Guth et al.[299,300] developed

2D and 3D structures for maximizing elastic and thermal constitutive properties. Others have

expanded on the work by exploring penalization for buckling [302], not limiting the design

volume to a predetermined size shape [301], or for minimum compliance applications [303,304].

Despite extensive research in the field, for many years one of the biggest challenges was the

ability to manufacture such structures.

With the recent increased interest in AM and improved fabrication resolution, optimized

metamaterial design is seeing renewed interest. The ground structure approach is well suited for

lattice structure design because it is a good representation of common lattice structures which

consist of struts and nodes, and can reduce the number of design variables compared to the

continuum-based approach.

Page 139: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

110

In this work, the ground structure was formed using truss elements which carry only axial

loads. The use of truss elements assumes the elements were connected by pin joints at the nodes.

The ground structure formulation is valid for linear geometric deformations. The design space

was assumed to be cubic with an equal number of nodes along each edge. The design variables

for the structure were the element cross-sectional areas. By adjusting the cross-sectional area of

each element, 𝐴𝑒, the optimization algorithm identified elements that were non-critical to

improving the objective function(s). The upper and lower bounds on the design variables were

𝐴𝑢𝑝𝑝𝑒𝑟 and 𝐴𝑙𝑜𝑤𝑒𝑟, respectively. As opposed to an element deletion technique for the ground

structure model, 𝐴𝑒 was allowed to approach a minimum, 𝐴𝑙𝑜𝑤𝑒𝑟, which was set to a small value

several orders of magnitude lower than the other element areas, yielding an insignificant impact

on the structure. In other words, as the cross-sectional area approaches 𝐴𝑙𝑜𝑤𝑒𝑟, that element has

negligible impact on the performance of the structure determined by the objective function for

optimization. As opposed to simply a sizing optimization, as 𝐴𝑒 approaches 𝐴𝑙𝑜𝑤𝑒𝑟, the topology

changes because elements become negligible and can later be discarded from the final structure.

3.2.2 Optimization Algorithms

Two optimization algorithms have been used for optimization of the ground structure

model in ALTO: fmincon, and gradient-based algorithm, and Borg, an evolutionary algorithm. A

brief description of how each algorithm works is explained in the sections that follow. In these

sections, general advantages and disadvantages of each algorithm are mentioned. For additional

information on the optimization algorithms and findings from this research see the following

sections:

• Section 1.1.6.1 – general comparison of gradient-based and evolutionary

algorithms,

• Sections 4.4 and 5.5 – discussion of fmincon and Borg for specific case studies,

Page 140: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

111

• Sections 7.1 and Section 7.2 – summary and primary conclusions.

3.2.2.1 fmincon: A Gradient-based Optimization Algorithm

MATLAB’s fmincon is a gradient-based nonlinear programming solver [358]. In cases

where the gradient of the system is not given, fmincon approximates the gradient numerically by

gradually adjusting each design variable to determine an appropriate search direction. The

convergence criteria for fmincon is based on when the objective values are no longer decreasing

to within the tolerance on the design variables. As with other gradient-based algorithms, fmincon

generally requires fewer function evaluations, has a clear convergence criterion, and is a faster

method than non-gradient based algorithms because it searches in a single direction based on the

gradient approximation. The downsides to gradient-based algorithms are that they are highly

dependent on the initial values for the design variables, easily become trapped in local-minima,

and can be challenging for considering multi-objective optimizations. For gradient-based

algorithms, multi-objective problems are generally reduced to a single objective using a weighted

sum of the objectives being considered. As a result, the optimization searches for a single solution

which is highly dependent on an appropriate selection of the weight values.

For a gradient-based optimization algorithm, such as fmincon, a description of the

optimization problem is given in Eqn. (3.1)-(3.7),

𝑀𝑖𝑛𝑖𝑚𝑖𝑧𝑒 𝑓(𝑊, 𝑃)

𝑆𝑢𝑏𝑗𝑒𝑐𝑡 𝑡𝑜 𝐴𝑒 − 𝐴𝑙𝑜𝑤𝑒𝑟 ≥ 0 (3.1)

𝐴𝑒 − 𝐴𝑢𝑝𝑝𝑒𝑟 ≤ 0 (3.2)

𝐴𝑒 = 𝐴𝑙𝑜𝑤𝑒𝑟 𝑓𝑜𝑟 𝐴𝑒 < 𝐴𝑟𝑒𝑙 (3.3)

𝑊ℎ𝑒𝑟𝑒 𝑓 = (𝑊)(𝑃) (3.4)

𝑊 =∑𝑐𝑖𝑚𝑖𝑤𝑖(𝑨, 𝑳), 𝑓𝑜𝑟 𝑖 = 1…𝑛𝑜𝑏𝑗

𝑛

1

(3.5)

Page 141: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

112

1 = ∑𝑐𝑖

𝑛

1

, 𝑓𝑜𝑟 𝑖 = 1…𝑛𝑜𝑏𝑗 (3.6)

𝑃 =∑𝑝𝑗(𝑨)

𝑛𝑝

𝑗

, 𝑓𝑜𝑟 𝑗 = 1…𝑛𝑝 (3.7)

where 𝐴𝑟𝑒𝑙 was a DfAM consideration called the relevant area limit discussed in Sections

3.2.3.1.5 and 0, the overall objective, 𝑓, was based on the combined objective value, 𝑊, and

combined penalty value, 𝑃. A weighted sum was used to calculate 𝑊, where 𝑤𝑖 were the

individual objective scores, 𝑐𝑖 were the corresponding weight values, and 𝑚𝑖 was either 1 or -1 to

minimize or maximize an objective, respectively. The number of objectives was 𝑛𝑜𝑏𝑗 and 𝑨, 𝑳

were vectors of the element cross-sectional areas and lengths, respectively. The combined penalty

value, 𝑃, was calculated by adding the individual penalty scores, 𝑝𝑖, for 𝑛𝑝 penalty functions.

Penalties are used to negatively impact the objective function score, making the design less

economical than non-penalized designs, without removing it from the solution space. Additional

information on the a DfAM penalty function for gradient-based optimization is given in Section

3.2.3.3.2.

3.2.2.2 Borg: A Multi-objective Evolutionary Algorithm

Borg is a multi-objective evolutionary algorithm (MOEA). Evolutionary algorithms are

heuristic approaches which have little or no information about the problem while iteratively

trying to improve the design. They evaluate a population of designs and then try to form new

designs by recombination and mutation of the design variables from the best performing designs.

Evolutionary algorithms tend to excel where gradient-based algorithms struggle: multi-objective

solutions, finding global minimums, and less dependence on a single initialization of the design

variables. They accomplish this by evaluating designs throughout the design space and taking the

designs that perform the best in each of the objectives into consideration. They maintain a set of

Page 142: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

113

good designs making it easier to search multiple directions and develop a set of Pareto-optimal

solutions. A Pareto-optimal solution is one for which no other design is better than the current

design in all objectives. The downside of evolutionary algorithms is that they are generally slow

because they are exploring multiple search directions and have no gradient-type information to

direct their search.

Borg was developed by Hadka and Reed [258] to address issues encountered with

traditional MOEA’s. In their analysis of MOEA’s, the pareto front is found to be typically only

developed in a very localized region of the solution space. This localized resolution of the

solution space is referred to as dominance resistance and ultimately leads to deterioration of the

pareto front. Deterioration is the case in which the best designs at the current iteration are actually

worse than designs simulated during previous iterations. Borg, based on 𝜖-MOEA [359],

incorporates an 𝜖-box dominance, 𝜖-progress, adaptive population sizing, and six different

recombination methods, where 𝜖 is a threshold or tolerance on the objective function values.

Through the incorporation of these features, Borg preserves the diversity of designs across the

pareto front and improves search efficiency compared to other MOEAs [258]. In Borg, the default

termination criterion is the number of design evaluations.

For an evolutionary algorithm , such as Borg, a description of the optimization problem is

given in Eqn. (3.8)-(3.12),

𝑀𝑖𝑛𝑖𝑚𝑖𝑧𝑒 (𝑊1,𝑊2, … ,𝑊𝑖) 𝑓𝑜𝑟 𝑖 = 1…𝑛𝑜𝑏𝑗

𝑆𝑢𝑏𝑗𝑒𝑐𝑡 𝑡𝑜 𝐴𝑒 − 𝐴𝑙𝑜𝑤𝑒𝑟 ≥ 0 (3.8)

𝐴𝑒 − 𝐴𝑢𝑝𝑝𝑒𝑟 ≤ 0 (3.9)

𝐴𝑒 = 𝐴𝑙𝑜𝑤𝑒𝑟 𝑓𝑜𝑟 𝐴𝑒 < 𝐴𝑟𝑒𝑙 (3.10)

𝑊ℎ𝑒𝑟𝑒 𝑊𝑖 = 𝑚𝑖(𝑠𝑖𝑃 + 𝑤𝑖(𝑨, 𝑳)) (3.11)

𝑃 =∑𝑝𝑗(𝑨)

𝑛𝑝

𝑗

, 𝑓𝑜𝑟 𝑗 = 1…𝑛𝑝 (3.12)

Page 143: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

114

where, 𝑊𝑖 was the penalized objective score, 𝑤𝑖 was the raw or individual objective score, 𝐴𝑟𝑒𝑙

was a DfAM consideration called the relevant area limit discussed in Sections 3.2.3.1.5 and 0.

Without the need to combine all the objectives into a single value using the weighted sum, the

penalty was applied individually to each objective and may be scaled with 𝑠𝑖 based on the how

heavily each objective should be penalized. As mentioned previously, 𝑚𝑖 was either 1 or -1 to

minimize or maximize an objective, respectively. The number of objectives was 𝑛𝑜𝑏𝑗 and 𝑨, 𝑳

were vectors of the element cross-sectional areas and lengths, respectively. Penalty functions

were utilized to negatively impact the objective function values, making the design less

economical than non-penalized designs, without removing the design from the solution space.

Additional information on penalty functions used with Borg is given in Section 3.2.3.3.1.

3.2.3 ALTO Objectives, Constraints, and Penalties

The purpose of the optimization was to minimize the objective function(s) subject to the

constraints. First, designs were evaluated for feasibility based on the optimization constraints.

Each design that satisfies all constraints was then evaluated for each objective function. Next,

each design was evaluated against the penalty functions. Penalties were distinguished from

constraints as they negatively impact the objective function values but do not render designs

infeasible and were applied after evaluation of the objective functions. As such, a design that was

penalized may be part of the solution space, however, designs which do not meet the constraints

were not. Constraints were applied to the design variables and restrict the range of possible values

for the design variables.

As the purpose of ALTO was an application-agnostic method for design and optimization

of unit cell topologies, a library of objective, constraint, and penalty functions were developed to

include a wide variety of applications. Based on the intended application, the user selects

Page 144: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

115

appropriate functions from the library to customize the optimization. In addition, the formulation

of ALTO with an evolutionary algorithm was designed to be robust for accepting additional

objective, constraint, or penalty functions. In general, the functions are described in the context of

L-PBF but the approach can be tailored to a variety of AM processes. A summary of the current

options within ALTO is presented in Table 3.1. Each of the objective, constraint, and penalty

functions are described in detail in the following subsections.

Table 3.1 Summary of ALTO objective, constraint, and penalty functions for unit cell generation and

their applicability to DfAM

ALTO Purpose DfAM

Objectives (Maximize or Minimize…)

Strain Energy Strain energy subject to a loading condition

Thermal Conductance Thermal conductance between two sets of nodes

Homogenization Relative error from a specified target constitutive matrix

Unit Cell Volume Total volume or relative error from a specified volume fraction Yes

Relevant Elements Number of elements within the unit cell to control complexity Yes

Powder Removability Factor (PRF) Proximity of elements with respect to other elements Yes

Optimization Constraints (Constrain design space for…)

Relevant Cross- Sectional Area Minimum feature size for DfAM Yes

Cross-Sectional Area Limits Upper and lower bounds on element area

Volume Fraction Allowable material volume for design space

Ground Structure Restrictions (Convert elements to non-design elements for…)

Build Angle Elements within a specified DfAM build angle range Yes

Element Length Elements within a specified length range

Surface Elements Elements on unit cell surface

Interconnectivity Nodes Connectivity to neighboring unit cells

Symmetry Element symmetry across any/all orthogonal midplanes

Penalties (Discourage solutions …)

Overlap Containing overlapping elements Yes

Connectivity Without connection to neighboring unit cells

Unsupported‒S Not self-supported for AM Yes

Unsupported‒N Not supported for AM by neighboring unit cells Yes

Page 145: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

116

3.2.3.1 Objectives

Six different objective functions have been included for calculation in ALTO and are

described in the following paragraphs. The objective function calculations are for strain energy,

thermal conductance, homogenization, volume, the number of relevant elements, and powder

removability.

3.2.3.1.1 Strain Energy

As with most topology optimization approaches, the structural compliance, or

proportionally the strain energy, was calculated and represents the stored elastic energy in the

structure. The purpose of the minimization of strain energy was to maximize the stiffness of the

structure subject to a specific loading condition. For structural loading, the stiffness matrix, �̃�,

was a symmetric matrix formed for the structure and then loads and boundary conditions were

applied to nodes on the structure. The quasi-static displacements of each node were computed

using Eqn. (3.13),

�̃�𝒖 = 𝑭 (3.13)

where 𝑭 and 𝒖 were vectors of size [𝑁𝑑𝑜𝑓𝑥1] and �̃� was a matrix of size [𝑁𝑑𝑜𝑓𝑥𝑁𝑑𝑜𝑓] where

𝑁𝑑𝑜𝑓 was the number of degrees of freedom of the system. The vector of externally applied

forces corresponding to each of the degrees of freedom for the nodes was represented by 𝑭. The

vector of displacements for each degree of freedom was represented by 𝒖. The global stiffness

matrix, �̃�, was a function of 𝐴𝑒, where 𝑒 = 1…𝑀 and 𝑀 was the total number of elements in the

system. In the ground structure model, 𝑭 and the Young’s modulus, 𝐸, are assumed constant and

the formulation is for linear elastic and small deformations with the nodes assumed to act like pin

joints as explained previously.

Page 146: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

117

The work done on the system by externally applied loads was proportional to the strain

energy, 𝑆𝐸. The strain energy was given by Eqn. (3.14). The strain energy objective function,

𝑤𝑆𝐸, is simplified by ignoring the 1

2 because it does not affect the optimization; see Eqn. (3.15).

SE =1

2𝒖�̃�𝒖 (3.14)

wSE = 𝒖𝑇�̃�𝒖 (3.15)

For gradient-based optimization algorithms, the strain energy was normalized to assist

with the weighted sum combination of multiple objectives. In this situation, the objective function

was modified to Eqn.(3.16),

𝑤𝑆𝐸̅̅̅̅ =𝑤𝑆𝐸𝑚𝑖𝑛 −𝑤𝑆𝐸

𝑤𝑆𝐸𝑚𝑖𝑛 −𝑤𝑆𝐸

𝑚𝑎𝑥 (3.16)

where 𝑤𝑆𝐸𝑚𝑖𝑛 and 𝑤𝑆𝐸

𝑚𝑎𝑥 are the strain energy calculated with all the element areas set to 𝐴𝑢𝑝𝑝𝑒𝑟

and 𝐴𝑙𝑜𝑤𝑒𝑟, respectively.

3.2.3.1.2 Thermal Conductance

The purpose of the thermal conductivity objective was to maximize or minimize the

thermal conductance between two sets of nodes within the structure. The thermal conductance

was defined in terms of conductance from one node, or set of nodes, 𝑁1, to another node, or set of

nodes, 𝑁2. Temperatures, 𝑇1 and 𝑇2, were assigned to the 𝑁1 and 𝑁2, respectively. The thermal

resistance of each truss element within the unit cell was calculated using Eqn. (3.17),

𝑅𝑒 =𝐿𝑒𝑘𝐴𝑒

(3.17)

Page 147: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

118

where 𝑅𝑒 was the element thermal resistance, 𝐿𝑒 was the element length, 𝐴𝑒 was the element

cross-sectional area, and 𝑘 was the material thermal conductivity.

Next, using an analog to Kirchhoff’s voltage and current laws for solving an electrical

circuit, a linear system of equations was developed to calculate the temperature at each node and

the heat flux through each element using 𝑅𝑒, 𝑇1, 𝑇2, and the connectivity of the elements in the

unit cell. By calculating the heat flux through each element, the effective heat flux, 𝑞𝑒𝑓𝑓, across

the unit cell was calculated from the total heat flux leaving 𝑁1 or entering 𝑁2. Finally, the thermal

conductance objective was calculated using Eqn. (3.18)

𝑤𝑇𝐶 =𝑞𝑒𝑓𝑓

𝑇1 − 𝑇2 (3.18)

where 𝑤𝑇𝐶 was the thermal conductance objective function from 𝑁1 to 𝑁2. The thermal

conductance may be maximized or minimized between 𝑁1 and 𝑁2, depending on the application;

see an example 2D schematic in Figure 3.2.

Figure 3.2 Example of traditional thermal circuits and analogy to the ground structure formulation

for calculating the unit cell thermal conductance

Page 148: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

119

In the thermal conductance calculation described above, two sets of nodes must be

defined and given specific temperatures. As a slight variation, a modified thermal conductance

formulation was available to maximize or minimize heat conduction from a set of nodes where

heat was supplied, to a heat sink node(s). In the modified formulation, heat was supplied to nodes

within the design space, similar to a current source within a circuit. As the input heat drives the

heat flux through the system, the average temperature at the heat input nodes, �̅�, determines the

effective thermal conductance. If the average nodal temperature was high, the thermal

conductance was low, because high nodal temperatures were required to drive the heat flux

through a high thermal resistance. On the other hand, if the average nodal temperature was low,

the thermal conductance was high because low temperatures were sufficient to drive the heat flux

through a low thermal resistance. The modified thermal conductance objective is shown in Eqn.

(3.19),

𝑤𝑀𝑇𝐶 =𝑞𝑒𝑓𝑓

�̅� (3.19)

where 𝑤𝑀𝑇𝐶 was the modified thermal conductance, 𝑞𝑒𝑓𝑓 was the effective heat flux through the

unit cell, and �̅� was the average node temperature of the nodes where heat was supplied.

For gradient-based optimization, the thermal conductance objective function was

normalized as shown in Eqn. (3.20),

𝑤𝑇𝐶̅̅̅̅ =𝑤𝑇𝐶𝑚𝑖𝑛 −𝑤𝑇𝐶

𝑤𝑇𝐶𝑚𝑖𝑛 −𝑤𝑇𝐶

𝑚𝑎𝑥 (3.20)

where 𝑤𝑇𝐶𝑚𝑖𝑛and 𝑤𝑇𝐶

𝑚𝑎𝑥e calculated with all the element areas set to 𝐴𝑙𝑜𝑤𝑒𝑟 and 𝐴𝑢𝑝𝑝𝑒𝑟,

respectively. Similarly, the objective function for 𝑤𝑀𝑇𝐶 may be modified to form the normalized

equivalent 𝑤𝑀𝑇𝐶̅̅ ̅̅ ̅̅ ̅.

Page 149: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

120

3.2.3.1.3 Homogenization

The purpose of the homogenization objective was to design a lattice structure with

tailored compliance or designer-specified mechanical properties. For a given material, the

relationship between stresses and strains was defined by the elastic constitutive matrix, �̃�.

Homogenization is a method for approximating �̃�, by calculating the homogenized elastic

constitutive matrix, �̃�𝐻, which is based on averaging the properties of a representative volume

element (RVE). Throughout this dissertation, the term “constitutive matrix” refers to the elastic

constitutive matrix. In the case of lattice structures, homogenization means taking the potentially

complex behavior of the unit cell topology and averaging it across an entire body or patterned

lattice structure.

The homogenization approach was based on a formulation developed by Sigmund and is

summarized as follows; for additional details; see references [297,298]. For 2D, the approach

requires three test strain cases to be applied to the unit cell (six for 3D). The test strain cases were

applied in each primary orthogonal direction (X and Y), and in the shear direction (XY). As the

2D case can be readily demonstrated in schematics, an example of the three test strain cases for

2D are shown in Figure 3.3. For expansion to 3D, three test strain cases were applied in the

primary orthogonal directions (X, Y, Z), followed by three test strain cases in the shear directions

(XY, YZ, XZ). The test strain cases were applied using periodic boundary conditions on the unit

cell to ensure edge compatibility with neighboring unit cells in the patterned lattice structure. The

periodic boundary conditions were based on work by van der Sluis et al. [360]. The approach is

valid for calculating the homogenized constitutive matrix for small deformations.

Page 150: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

121

Figure 3.3 Schematic for three test strain cases to an RVE for calculation of the homogenized

constitutive matrix, adapted from [298]

Using the standard finite element method (see Section 3.2.3.1.1) to solve each of the test

strain cases, the resulting global displacement vector, 𝒖, was solved. Based on the deformations

and element stiffnesses, averaged mechanical properties were calculated for the unit cell from the

element mutual energies using Eqns. (3.21)-(3.22),

𝐸𝑘𝑙𝐻 =∑𝑄𝑘𝑙

𝑒

𝑀

𝑒=1

(3.21)

𝑄𝑘𝑙𝑒 =

1

𝑌(𝒅0𝑒

𝑘 − 𝒅𝑒𝑘)𝑇�̃�𝑒(𝒅0𝑒

𝑙 − 𝒅𝑒𝑙 ), 𝑓𝑜𝑟 {

𝑒 = 1,… ,𝑀 2𝐷: 𝑘, 𝑙 = 1,2,3 3𝐷: 𝑘, 𝑙 = 1,… ,6

} (3.22)

where �̃�𝐻 was the homogenized constitutive matrix, �̃�𝑒 was the element mutual energy, 𝒅0𝑒𝑘 was

the induced element displacement vector for each test strain case, 𝒅𝑒𝑘 was the applied element

displacement vector for each test strain case, �̃�𝑒 was the element stiffness matrix (4x4 for 2D, or

6x6 for 3D), 𝑀 was the number of elements, and 𝑌 was the characteristic length of the RVE. In

�̃�𝐻, the terms 𝐸11𝐻 , 𝐸22

𝐻 , 𝐸33𝐻 are the effective elastic modulus values for the homogenized

material, heretofore referenced as “effective modulus” values.

After calculating the constitutive matrix, the homogenization objective function was

calculated using Eqn. (3.23),

X

Y

Page 151: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

122

𝑤𝐻 =∑|𝐸𝑘𝑙

∗ − 𝐸𝑘𝑙𝐻 |

𝐸𝑘𝑙∗ , 𝑓𝑜𝑟 {

2𝐷: 𝑘, 𝑙 = 1,2,3 3𝐷: 𝑘, 𝑙 = 1,… ,6

} (3.23)

where �̃�∗ was the desired or target homogenized constitutive matrix. This objective function,

Eqn. (3.23), represented the total relative error between the target and calculated homogenized

constitutive matrices.

3.2.3.1.4 Unit Cell Volume

The purpose of the unit cell volume objective was to match a target unit cell volume

fraction or minimize/maximize the total unit cell volume. For matching a target volume fraction,

𝑉𝑓∗, the objective function value was the relative error between the target and calculated volume

fraction; see Eqn. (3.24),

𝑤𝑉𝐹 =|𝑉𝑓

∗ −𝑨𝑇𝑳𝑉𝑢𝑐

|

𝑉𝑓∗

(3.24)

where 𝑤𝑉𝐹 was the volume fraction objective function and 𝑉𝑢𝑐 was the total unit cell volume. On

the other hand, instead of using a target volume fraction, the total volume of material within the

unit cell can be minimized or maximized; see the volume objective function, 𝑤𝑉, in Eqn. (3.25).

In both methods, the volume calculation does not account for material overlap from neighboring

or intersecting elements.

𝑤𝑉 = 𝑨𝑇𝑳 (3.25)

The unit cell volume objective was indirectly used in the context of DfAM because the

volume of material within the unit cell can affect manufacturability. While in general the

complexity or amount of material is not a major concern for AM, the length scale for designing

unit cell topology quickly approaches the minimum feature size, making it difficult to

Page 152: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

123

individually print each intended strut. For example, high volume fraction unit cells are more

likely to lead to trapped powder than low volume fraction unit cells [245]. In addition,

minimizing or matching a specific volume fraction indirectly reduces complexity by encouraging

designs with low volume. In the subsequent section, a more direct approach than the volume

fraction objective for reducing complexity is addressed.

3.2.3.1.5 Number of Relevant Elements

In DfAM, component design complexity or intricacy is often less of a consideration than

with subtractive manufacturing processes because, arguably, design complexity is free. However,

when considering DfAM for unit cells, the unit cell intricacy can play a role in the

manufacturability because the unit cells are small and the elements or features of the unit cell are

often at or near the minimum feature resolution of the manufacturing process. Depending on the

laser scan strategy of the lattice structures, added complexity can also increase build time due to a

large number of short infill and contour laser passes. Asadpoure et al. [294] address the same

issue using a smooth and differentiable objective function for gradient-based optimization

approaches.

The purpose of the number of relevant elements objective was to reduce the design

complexity, encouraging fewer elements or struts within the unit cell solution. In this text,

“relevant element(s)”, means elements that have a cross-sectional area larger than 𝐴𝑟𝑒𝑙. Based on

the relevant cross-sectional area limit, 𝐴𝑟𝑒𝑙 (see Section 3.2.3.2.1), and by applying Eqn. (3.3),

any element cross-sectional areas that fell below 𝐴𝑟𝑒𝑙 were temporarily set to 𝐴𝑙𝑜𝑤𝑒𝑟 for the

current iteration. The relevant elements objective, 𝑤𝑅𝐸, was simply the current number of

elements with a cross-sectional area larger than 𝐴𝑟𝑒𝑙; see Eqn. (3.26).

𝑤𝑅𝐸 = 𝑐𝑜𝑢𝑛𝑡(𝐴𝑒 > 𝐴𝑟𝑒𝑙) (3.26)

Page 153: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

124

For DfAM, reducing the number of relevant elements reduced the complexity of the unit

cell. Unit cells that have a large number of struts are more likely to blend or mix the struts than

unit cells with few struts due to the process tolerances and minimum feature size. As a result, the

as-printed lattices will not match the as-designed model nor perform as predicted or intended. In

addition, more struts will generally lead to small pores and cavities where powder can become

partially sintered or even sealed off sections of the unit cell, making powder removal impossible.

3.2.3.1.6 Powder Removability

The powder removability objective function was for evaluation of the risk of trapping

powder within the unit cell. The purpose of the objective was to encourage unit cells with large

pores and spacing between relevant elements. Small pores or tight spacing between elements can

trap unsintered powder in L-PBF components [244,245]. A simple schematic for how powder can

become trapped is shown in Figure 3.4, where, as the build progresses, cavities may become

entirely or partially sealed off. Whether AM lattices are applied to aerospace components where

trapped powder adds weight to the component or medical components in which any unsintered

powder that eventually escapes could be toxic, it is critical to be able to efficiently remove all

unsintered powder from lattice structures. Aside from trapping powder, small pores also increase

the post-processing time to remove powder from a maze of intricate channels.

Page 154: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

125

Figure 3.4 Schematic showing how powder can become trapped in internal cavities for powder-based

AM processes.

In this dissertation, a new metric to measure powder removability from strut and node

type lattices is proposed. To the authors’ best knowledge, this is the first objective function or

evaluation related to powder removability for AM lattice structures. The powder removability

factor (PRF) was based on evaluation of the spacing between elements. The midpoint distance

between each element to every other element in the unit cell was calculated based on the initial

full ground structure, where �̃�𝑒𝑒 was the element-to-element distance matrix. Within each

position of the matrix, 𝐷𝑖𝑗𝑒𝑒 was the distance from midpoint to midpoint between element 𝑖 and

element 𝑗. During the PRF objective evaluation of each design, the radius of each element was

subtracted; see Eqn. (3.27),

𝑃𝑖𝑗 = 𝐷𝑖𝑗

𝑒𝑒 − 𝑟𝑖𝑒 − 𝑟𝑗

𝑒 , 𝑓𝑜𝑟 {𝑖 = 1,… ,𝑀𝑗 = 1,… ,𝑀

} (3.27)

where �̃� was the pore distance matrix and 𝑟𝑖𝑒, 𝑟𝑗

𝑒 were the radii of the i-th and j-th elements. An

example of the pore size measurement is shown in Figure 3.5.

Page 155: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

126

From the resulting matrix containing all the distance measurements for relevant elements

(𝐴𝑒 > 𝐴𝑟𝑒𝑙), the PRF objective score was the number of instances where the distance fell below a

given threshold. Eqn. (3.28) shows the raw objective function for the PRF objective (𝑤𝑃𝑅𝐹),

𝑤𝑃𝑅𝐹 = 𝑐𝑜𝑢𝑛𝑡(𝑃𝑖𝑗 < 𝑝∗) (3.28)

where 𝑝∗ is the pore threshold.

The pore threshold value, 𝑝∗, may vary based on the process and intent of the objective.

For example, to simply improve powder removability during post processing, 𝑝∗ is chosen to

encourage pores larger than the specified threshold, improving flowability of the powder out of

the lattice. On the other hand, to specifically discourage designs that are likely to have closed

pores due to overmelt or dross, excess powder particles sintered to the downskin surface, 𝑝∗ may

be chosen based on the minimum gap size for which powder may escape. In this second case 𝑝∗ is

directly dependent on the AM process, powder size distribution, laser process parameters, etc.

Figure 3.5 Ground structure-based pore size measurement for the PRF. On the left part of the image,

an example of the element-to-element distance measurement is made midpoint to midpoint and then

the radius of each element is subtracted to obtain the pore size for the pore distance matrix. A few

examples of some of the pore measurements are shown on the right part of the image. In the pore

matrix, each element is measured relative to every other element.

Page 156: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

127

3.2.3.2 Optimization Constraints and Ground Structure Restrictions

Eight different constraints and restrictions can be applied to the ground structure and

optimization problem to reduce the number of design variables and constrain design space. Three

were optimization constraints that control the relevant cross-sectional area, cross-sectional area

limits, and volume fraction. The remaining five are restrictions on the ground structure model

configuration for DfAM and are not enforced as constraint equations in the optimization problem.

Each of these optimization constraints and ground structure restrictions are described in the

following sections.

3.2.3.2.1 Relevant Cross-Sectional Area

The relevant cross-sectional area constraint is for elements that were deemed not

manufacturable because they fell below the minimum feature size. The constraint was based on

the relevant cross-sectional area limit, 𝐴𝑟𝑒𝑙, which defines how large an element should be to be

printable and structurally sound. Eqn. (3.3) and Eqn. (3.10) define the relevant cross-sectional

area constraint. Elements with a cross-sectional area less than 𝐴𝑟𝑒𝑙 are temporarily set to the

lower limit, 𝐴𝑙𝑜𝑤𝑒𝑟, for the current iteration and remain in the optimization as design variables.

𝐴𝑟𝑒𝑙 was based on the AM process-specific limitations. For example, in L-PBF, the

manufacturable spot size is typically close to 150-200 microns in diameter. From this spot size,

the generally accepted minimum feature size for L-PBF printers is a diameter of 300-500 microns

so that more than a single point can be used to represent the contour or infill. In this work, where

all element cross sections were assumed to be circular, 𝐴𝑟𝑒𝑙 was defined by the relevant diameter

limit, 𝑑𝑟𝑒𝑙, as shown in Eqn. (3.29).

𝐴𝑟𝑒𝑙 =

𝜋𝑑𝑟𝑒𝑙2

4 (3.29)

Page 157: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

128

3.2.3.2.2 Cross-Sectional Area Limits

The cross-sectional area of each element in the system was constrained by upper and

lower bounds, 𝐴𝑢𝑝𝑝𝑒𝑟 and 𝐴𝑙𝑜𝑤𝑒𝑟. The purpose of the lower limit was to serve as a sufficiently

low bound that any element near or set to this value had a negligible contribution on the system

and objective function scores. The purpose of the upper limit was to help determine the allowable

topologies based on the largest feature size of the unit cell. The upper bound was chosen based on

the desired diameter upper limit, 𝑑𝑢𝑝𝑝𝑒𝑟. The lower diameter limit, 𝑑𝑙𝑜𝑤𝑒𝑟, was set such that

𝐴𝑙𝑜𝑤𝑒𝑟 ≪ 𝐴𝑢𝑝𝑝𝑒𝑟. The effect of 𝑑𝑢𝑝𝑝𝑒𝑟 depended on the objective of the optimization. For

example, for high stiffness unit cells, increasing 𝑑𝑢𝑝𝑝𝑒𝑟 may enable unit cell topologies with

thicker and fewer elements. In general, the relevant diameter limits were chosen such that 𝐴𝑙𝑜𝑤𝑒𝑟

was five or more orders of magnitude smaller than 𝐴𝑢𝑝𝑝𝑒𝑟. The exact relevant diameter limits

were chosen based on the optimization problem and were specified for each optimization case

study.

3.2.3.2.3 Volume Fraction

In cases where the volume was not used as an objective, the volume fraction can be a

constraint that limits the total volume that can be included in the unit cell. The purpose was to

restrict the allowable material volume, especially in cases where other objectives encourage high

volume fraction unit cells. The constraint was formulated using Eqn. (3.30),

𝑉𝑓∗𝑉𝑢𝑐 ≥ 𝑨

𝑇𝑳 (3.30)

where 𝑉𝑓∗ was the target volume fraction, and 𝑉𝑢𝑐 was the total volume of the unit cell. A common

approach to GSTO is also to use the total possible volume, 𝑉𝑚𝑎𝑥, which is the combined volume

of all elements at 𝐴𝑢𝑝𝑝𝑒𝑟 instead of the total enclosed volume of the unit cell.

Page 158: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

129

3.2.3.2.4 Ground Structure Restrictions

For a full ground structure, each node is connected to every other node by an element.

While a full ground structure lends the most design freedom to the optimization for a given set of

nodes, the number of elements or design variables, 𝑀, increases based on Eqn. (3.31),

𝑀 =

𝑁𝑡(𝑁𝑡 − 1)

2 (3.31)

where 𝑁𝑡 was the total number of nodes in the ground structure. As such, the number of design

variables can quickly become unwieldy for optimization. In addition, a full ground structure was

highly redundant due to overlapping elements. For example, for a unit cell defined by a 5x5x5

array of nodes, the total number of nodes is 125, leading to 7750 design variables in a full ground

structure. To reduce the number of design variables, restrictions were placed on the ground

structure configuration, permanently converting the elements to non-design elements in the

optimization problem.

There were five restrictions that could be applied to reduce the ground structure

configuration: build angle, element length, surface elements, interconnectivity nodes, and

symmetry. In each of these restrictions, elements from the initial ground structure that did not

meet the criteria were set as non-design elements. Non-design elements were present in the

ground structure model but were not design variables in the optimization. The cross-sectional

areas of non-design elements were either permanently set to a fixed value (i.e., 𝐴𝑒 = 𝐴𝑙𝑜𝑤𝑒𝑟) or

linked to another element with symmetry. As described in Section 3.2.1, as opposed to removal

from the ground structure model, non-design elements maintain model continuity and prevent

singularity of the global stiffness matrix.

In AM, the build angle is defined as the angle between horizontal build plate and the

component surface. Due to the nature of AM, built layer-by-layer, low build angles can cause

Page 159: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

130

manufacturability issues in some AM processes. The purpose of the build angle restriction was to

improve manufacturability of the unit cell by converting elements that would be difficult to build

to void non-design elements. Eqn. (3.32) defines the restriction,

𝐴𝑒 = 𝐴𝑙𝑜𝑤𝑒𝑟 𝑓𝑜𝑟 𝜃𝑙𝑜𝑤𝑒𝑟 < 𝜃𝑒 < 𝜃𝑢𝑝𝑝𝑒𝑟 (3.32)

where 𝜃𝑢𝑝𝑝𝑒𝑟 and 𝜃𝑙𝑜𝑤𝑒𝑟 are the upper and lower limits on the build angle and 𝜃𝑒 is the element

build angle, measured as smallest angle between the build plate and element direction vector.

The upper and lower limits of the build angle were chosen based on the AM process.

Generally, for L-PBF, the upper limit was set between 30-45 degrees. As processing and

manufacturing parameters continue to improve, small support-free build angles can be achieved

with reasonable surface roughness. In this work, the lower limit was often chosen as 1 degree,

allowing horizontal elements (𝜃𝑒 = 0) because at the length scale of lattice structures it is

frequently possible to span between two vertical supports. As shown in Figure 3.6, there is an

intermediate build angle range which is difficult to build due to the low build angle.

Figure 3.6 For lattice structures, horizontal elements are allowed to span between two vertical

supports because of the generally small length-scale. However, there is an intermediate range for

which the build angle is too low to accurately build.

Page 160: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

131

As reported in the author’s previous work, based on our preliminary studies of lattice

structures, the metal powder bed fusion process can create horizontal members for bridging [361].

Bridging refers to horizontal members connected at each end, forming a bridge between two

members. The term bridging is common in fused filament fabrication (FFF), where the filament is

stretched horizontally across two supports. Figure 3.7 shows several examples of horizontal

elements printed in lattice structures built with metal powder bed fusion, including 3x3x3 arrays

of 4 mm and 7 mm octet-truss unit cells, and 4 mm and 6 mm cubic unit cells. Note that the

horizontal members do print, however, for high element aspect ratios (length/diameter), or in the

case of cantilevers, distortion would occur, resulting in poor print quality and increased likelihood

of connection failures in the lattice.

Figure 3.7 Horizontal lattice members are shown in two lattice examples, 3x3x3 arrays of octet-truss

unit cells (4 mm, 7 mm) and cubic unit cells (4 mm, 6 mm). These lattices were printed in Ti64 on a

3D Systems ProX DMP 320. Image adapted from [361].

The element length restriction converted elements whose length fell within a specified

range to void non-design elements. The purpose of the restriction was to find only designs with

elements outside of the specified length range. Eqn. (3.33) defines the element length restriction,

𝐴𝑒 = 𝐴𝑙𝑜𝑤𝑒𝑟 𝑓𝑜𝑟 𝐿𝑙𝑜𝑤𝑒𝑟 < 𝐿𝑒 < 𝐿𝑢𝑝𝑝𝑒𝑟 (3.33)

Page 161: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

132

where 𝐿𝑢𝑝𝑝𝑒𝑟 and 𝐿𝑙𝑜𝑤𝑒𝑟 were the upper and lower limit, respectively, on element length, 𝐿𝑒.

Any element(s) within this length range were set to 𝐴𝑙𝑜𝑤𝑒𝑟 in the ground structure model and

became void non-design elements.

With the cubic design space, the surface elements restriction converted elements that lie

on the cubic faces of the unit cell to void non-design elements and the cross-sectional area is set

to the lower limit. The purpose of setting the cross-sectional area of the surface elements from the

unit cell to the lower limit was to reduce the impact on neighboring structures or connectivity to

other unit cells. This was accomplished by examining the starting and ending node locations of

each element to determine which elements lie solely on the cube faces.

Interconnectivity nodes were defined as nodes that lie on the surface of the unit cell cube.

These nodes determine the connectivity to neighboring unit cells or component structures. The

purpose of the interconnectivity nodes restriction was to limit how the unit cell connects with

itself or neighboring structures. For a given set of nodes, the cross-sectional area of any element

connected to one of these nodes was set to the lower limit and the element was converted to a

void non-design element. Using this restriction did not necessarily set the cross-sectional area of

all surface elements to the lower limit. Surface elements connected to two nodes that are not part

of the interconnectivity nodes restriction would remain as design variables in the optimization.

Finally, the symmetry restriction links the cross-sectional areas of symmetric elements to

a single design variable and converts the elements to non-design elements. The purpose of the

symmetry restriction was to reduce the number of design variables. Symmetry was applied based

on the three mid-planes (X, Y, Z) of the cubic design domain, individually or in any combination

of the three, resulting in seven distinct symmetry restriction options (X, Y, Z, XY, XZ, YZ,

XYZ).

Page 162: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

133

3.2.3.3 Penalties

For penalty functions, two different approaches were taken based on the type of

algorithm. To differentiate penalty functions from constraints, constraints were evaluated prior to

evaluating the objective functions of a unit cell design. Constraints restricted the design variable

values and if the design violated any of the constraints, the unit cell design was removed from the

solution space. In contrast, penalty functions were evaluated after the objective function(s) and

may negatively impact the objective function score, but the unit cell design may still be included

in the solution space.

3.2.3.3.1 Evolutionary Algorithm Penalty Functions

For the evolutionary algorithm, a set of penalties have been created. For each penalty, if

the current design being evaluated does not meet the criteria of the penalty, a flag was triggered.

As shown in Eqn. (3.12), the sum of the total number of flags was calculated as 𝑃. Using 𝑠𝑖 the

penalty 𝑃 can be scaled based on the relative order of magnitude of the corresponding objective

function or by the magnitude of the penalty impact. Next the penalty value, 𝑠𝑖𝑃, was added to

each raw objective score, 𝑤𝑖, and then multiplied by 𝑚𝑖 to determine whether an objective should

be minimized (𝑚𝑖 = 1) or maximized (𝑚𝑖 = −1); see Eqn. (3.11).

Four types of penalties may be applied to the unit cell optimization to discourage

infeasible or undesirable unit cell solutions without removing them from the solution space. The

four penalties were overlap, connectivity, and two options for verifying AM support,

unsupported‒S or unsupported-N. For each penalty, the unit cell was evaluated using a short

computational function to check for any violations. If any violations were detected during the

check, the penalty being checked was flagged. As shown in Eqn. (3.12), the penalty, 𝑃, was the

sum of the flags raised during the evaluation of the unit cell.

Page 163: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

134

3.2.3.3.1.1 Overlap Penalty

With a full ground structure, where an element connects each node to every other node in

the system, there were numerous overlapping or intersecting elements due to the highly redundant

elements. Mathematically, elements can overlap partially or entirely with no problem, but

physically two separate elements cannot occupy the same physical space. Similarly, elements can

intersect at locations that were not their endpoints mathematically, but physically these were not

desirable. The purpose of the overlap penalty was to discourage a unit cell design that had

overlapping or intersecting elements at non-end points of the elements.

Using the line vector that defined the position and direction of each element, the overlap

function determined which line vectors overlapped or intersected at non-end points of the

element. Overlapping or intersecting line vectors for elements were stored in a binary 𝑀x𝑀

matrix, where 𝑀 was the total number of elements in the full ground structure. To increase speed,

the overlap matrix was computed once for the ground structure as opposed to every iteration,

meaning that the current radius of each overlapping element was not considered.

During each iteration, the overlap matrix was reduced to only the relevant elements. If

any relevant elements for the iteration overlap or intersect according to the overlap matrix, a flag

was raised for the overlap penalty. Evaluation of the overlap penalty is shown in Eqn. (3.34)

𝑝𝑜(𝑨) = {

0, �̃� = �̃�

1, �̃� ≠ �̃� (3.34)

where 𝑝𝑜 is the overlap penalty, �̃� is the overlap matrix, and �̃� is the null matrix.

3.2.3.3.1.2 Connectivity Penalty

The connectivity penalty function was designed to prevent free hanging elements that

would not connect to any neighboring unit cells. For each patterned unit cell, there are 26

neighboring unit cells that interact with it. The simplest way to visualize this is to consider a

Page 164: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

135

3x3x3 array of unit cells, for which there is a single central cell that is connected to the other 26

unit cells surrounding it.

The connectivity computational function first determined node-neighbor pairs, or nodes

that were shared by neighboring unit cells (i.e. face-centered nodes on opposing faces). Next, it

determined which nodes were used by relevant elements within the unit cell. Finally, it checked

to ensure that if any nodes were used by relevant elements, at least one of their node-neighbor

pairs were also used by other relevant elements within the unit cell, creating a binary connectivity

vector marking whether a violation had occurred. The purpose of the connectivity penalty is to

ensure that the unit cell will connect to neighboring unit cells; see Eqn. (3.35),

𝑝𝐶(𝑨) = {0,𝑩 = 𝑶1,𝑩 ≠ 𝑶

(3.35)

where 𝑝𝑐 is the connectivity penalty, 𝑩 is the connectivity vector, and 𝑶 is the null vector. If

connectivity was flagged, that signified that there were relevant elements that would be free

hanging and not connected to relevant elements in the neighboring unit cells when patterned. In

optimization problems where homogenization was not the optimization objective, the

connectivity penalty could be used to ensure that relevant elements connect to other relevant

elements in neighboring unit cells.

3.2.3.3.1.3 Unsupported‒S Penalty

A major manufacturing limitation that is unique to most AM processes is the need for

support material. In L-PBF, support material is used to prevent distortion and ensure there is a

surface present on which to build the layer. For lattice structures fabricated with AM, the

presence of support material is generally infeasible because it is inaccessible and cannot be

removed from the intricate cavities and geometries that are created. The purpose of the

Page 165: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

136

unsupported‒S penalty (unsupported by self) was to ensure the unit cell can be printed without

support material.

With this penalty, the unit cell was checked to ensure that each relevant element was self-

supported by other elements within the unit cell. Based on the relevant elements within the

structure, each node was determined to be supported or unsupported. Layer by layer for nodes in

the XY planes of the ground structure model, each node is determined to be supported or

unsupported based on the relevant elements of the current design. A supported node is a node that

is supported by relevant elements below it and upon which other relevant elements could feasibly

build. In the current formulation, there is no max overhang or max allowable bridging distance

within a unit cell topology. All horizontal cantilevered elements (connected to a single supported

node) are considered unsupported and would trigger the penalty. All horizontal elements that

bridge between two supported nodes are considered supported and not penalized. The assumption

is that elements feasibly can bridge horizontally in the XY plane across two supported nodes

within the unit cell topology. As such, bridge distance is not a parameter for unsupported

penalties and acts under. If a relevant element was connected to an unsupported node, a violation

was marked in a binary unsupported elements vector. In addition, for a self-supported unit cell,

any nodes on the bottom cube face that were connected to relevant elements should have had their

mirrored opposite nodes also connected to relevant elements on the upper plane to ensure that for

a vertical column of patterned unit cells it continued to be self-supported. Eqn. (3.36) shows the

unsupported‒S penalty,

𝑝𝑈𝑆(𝑨) = {0, 𝑺 = 𝑶1, 𝑺 ≠ 𝑶

(3.36)

where 𝑝𝑈𝑆 is the unsupported‒S penalty, 𝑺 is the unsupported elements vector, and 𝑶 is the null

vector. If the unsupported‒S penalty was flagged, it signified that there were relevant elements

Page 166: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

137

building vertically upward from nodes that were unsupported. These were elements that would

not build properly without added support material.

3.2.3.3.1.4 Unsupported‒N Penalty

There were cases for which a relevant element could be supported by its neighboring unit

cells but it would not qualify as a self-supported unit cell on its own. The purpose of the

unsupported‒N penalty (unsupported by neighbors) was to ensure that the unit cell can be built

without added support, as long as when it was patterned it had other unit cells built around it in

the same plane.

Similar to the unsupported‒S penalty, the unsupported‒N determined which nodes were

supported by relevant elements in the unit cell. Next, considering node-neighbor pairs described

in Section 3.2.3.3.1.2, the list of supported nodes could be augmented by nodes shared by

neighboring unit cells. As such, the structures that pass the unsupported‒S penalty check would

be a subset of any structures that passed the unsupported‒N penalty. Similar rules for horizontal

cantilevered or bridging elements apply as described for the unsupported-S penalty. Finally, if a

relevant element was connected to an unsupported node, a violation was marked in a binary

unsupported elements vector. Eqn. (3.37) shows the unsupported‒N penalty,

𝑝𝑈𝑁(𝑨) = {

0, 𝑺𝑁 = 𝑶1, 𝑺𝑁 ≠ 𝑶

(3.37)

where 𝑝𝑈𝑁 is the unsupported‒N penalty, 𝑺𝑁 is the unsupported elements vector considering

neighbored support nodes, and 𝑶 is the null vector.

Similar to the connectivity penalty, unsupported‒N considers relevant elements that

would be connected to the neighboring unit cells that could provide support for what would

otherwise be free hanging structures. The unsupported‒N penalty assumes an infinite array of unit

cells in the build plane and does not consider unit cells on the boundary of a finite array.

Page 167: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

138

To aid with visualization of the connectivity, unsupported-S, and unsupported‒N

penalties, a few 2D examples are shown in Table 3.2. As any structures that pass the

unsupported‒S penalty are a subset of the unsupported‒N penalty, it is not possible to have a

structure pass unsupported‒S but not unsupported‒N.

Table 3.2 Examples of 2D unit cells for comparing pass/fail criteria of different penalty functions.

ALTO Penalties

Unit Cell Lattice Connectivity Unsupported‒S Unsupported‒N

Fail Fail Fail

Pass Fail Fail

Pass Fail Pass

Fail Pass Pass

Pass Pass Pass

Figure 3.8 shows a 2D example of determining supported and unsupported nodes for the

unsupported‒S and unsupported‒N penalties. In this example, the unit cell topology would pass

the unsupported‒N penalty but not the unsupported‒S penalty (row 3 of Table 3.2). There are

three rows of elements in the XY plane and non-relevant elements are not shown. First, the

relevant elements that start at row 1 determine which nodes are supported in rows 2 and 3. The

supported nodes are added to a list of supported nodes. Next, node-neighbor pairs are considered

and added to the list. Then, the next set of relevant elements starting from supported nodes in row

2 determines additional nodes supported in row 3. The top row determines which nodes on the

Page 168: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

139

bottom row are supported. Horizontal elements that bridge supported nodes are considered

feasible. After the list of all supported nodes in the unit cell has been compiled, it is compared

against a list of all nodes associated with relevant elements. If there are any nodes that are not in

the supported list, then the unsupported‒N penalty is triggered.

The unsupported‒S penalty check is performed similarly but without the node-neighbor

pairs check for which a node supported by a neighboring unit cell would augment the list of

supported nodes. In this example, node 4 would not have been added for the unsupported‒S

penalty and in the final check would have identified two unsupported elements that connect to

node 4. This basic example is shown in 2D for simplicity of explaining the process but the

penalties are extended to 3D for ALTO.

Page 169: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

140

Figure 3.8 Basic 2D example for determining the supported nodes for the unsupported‒N penalty.

This example comes from row three of Table 3.2 which passes the unsupported‒N penalty but would

fail the unsupported‒S penalty. The unsupported‒S penalty follows a similar process but the node-

neighbor pairs would not be considered meaning that node 4 is not self-supported, but instead

supported by a neighboring unit cell. Horizontal elements that bridge supported nodes are

considered feasible.

3.2.3.3.2 Gradient-based Algorithm Penalty

For the gradient-based optimization algorithm, a continuous penalty function was

formulated. The purpose of the penalty function was to reduce the complexity of the final

Page 170: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

141

topology by penalizing unit cell topologies with a higher number of relevant elements than other

topologies. By reducing the number of elements in a unit cell, the unit cell becomes more

manufacturable because there were fewer elements that can overlap and solidify together during

printing. The penalty function was defined by Eqns. (3.38)-(3.42),

𝑝𝑠𝑒 = 𝛼3 (

𝐴𝑟𝑒𝑙𝐴𝑒

) + 𝛼4 , 𝑓𝑜𝑟 𝐴𝑒 < 𝛼2𝐴𝑟𝑒𝑙 (3.38)

𝑝𝑠𝑒 = 1, 𝑓𝑜𝑟 𝐴𝑒 ≥ 𝛼2𝐴𝑟𝑒𝑙 (3.39)

𝛼3 =

𝛼2(𝛼1 − 1)

𝛼2 − 1 (3.40)

𝛼4 =𝛼2 − 𝛼1𝛼2 − 1

(3.41)

𝑝𝑖 = (𝑝𝑠1 ∗ 𝑝𝑠2 ∗ …∗ 𝑝𝑠𝑒)𝛾 , 𝑓𝑜𝑟 𝑒 = 1…𝑛𝑅𝐸 (3.42)

where 𝑛𝑅𝐸 was the number of relevant elements, 𝑝𝑠𝑒 was the penalty associated with each

relevant element, 𝛾 was used to change the importance of the penalty, and (𝛼1, 𝛼2, 𝛼3, 𝛼4) were

parameters to define the penalty function values and range.

By using the penalty function, elements with an area close to 𝐴𝑟𝑒𝑙 become less

economical in improving the objective function than elements not close to 𝐴𝑟𝑒𝑙, encouraging few

elements with large cross-sectional areas as opposed to many elements with small cross-sectional

areas. An example plot of the penalty function is shown in Figure 3.9. In this case (𝛼1, 𝛼2, 𝛼3,

𝛼4) were set to (2, 100,100

99,98

99), respectively. The result is that for elements at 𝐴𝑟𝑒𝑙, 𝑝𝑠𝑒 = 2. The

value of 𝑝𝑠𝑒 decreases as 𝐴𝑒 increases until 𝐴𝑒 = 𝛼2𝐴𝑟𝑒𝑙, at which point 𝑝𝑠𝑒 = 1.

Page 171: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

142

Figure 3.9 Graph of penalty function that reduces complexity by encouraging elements with a cross-

sectional area larger than the relevant cross-sectional area limit.

3.2.3.4 ALTO Summary and DfAM Considerations

As a major aspect of ALTO was to enable unit cell DfAM, the DfAM considerations are

reiterated in this section as well as noted in Table 3.1. By selectively choosing different

combinations of optimization components from Table 3.1, ALTO enables researchers and

designers to optimize lattice structures for specific applications and take advantage of the design

freedom offered by AM.

The primary AM-related criteria considered in ALTO were minimum feature size, design

complexity, and support material. The purpose of these DfAM considerations was to improve

manufacturability of the model’s optimal solution and reduce the need for design interpretation

and manipulation prior to fabrication the physical solution.

1) The minimum feature size was controlled by 𝐴𝑟𝑒𝑙. Elements with an area less than 𝐴𝑟𝑒𝑙

were considered infeasible because they would be difficult to reliably manufacture.

Page 172: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

143

2) Design complexity was reduced by including objective functions such unit cell volume

and number of relevant elements, or through an overlap penalty. Minimization of the unit

cell volume or number of relevant elements encouraged less material volume and

elements. In both cases, fewer elements or less material reduced the risk of overlapping

elements and blending or mixing of elements printed too close to each other. The overlap

penalty reduced complexity and time required for design interpretation.

3) Designs that would require support material were discouraged or eliminated through the

build angle restriction and the unsupported‒S or unsupported‒N penalties.

3.2.4 ALTO Validation

Three types of benchmark problems were used for validation of different components of

ALTO. The benchmark types were minimization of strain energy, maximization of thermal

conductance, and homogenization. In optimization, it is difficult to prove that the result is not a

false positive, a result that is not truly optimal. This would occur if the optimization is not able to

find the global minimum. The purpose of the validating ALTO against these benchmark problems

is to show that the ground structure formulation and optimization are functioning as expected and

that ALTO is able to replicate similar results to those found in literature for standard topology

optimization problems.

3.2.4.1 Minimization of Strain Energy

Minimization of compliance benchmarks were test problems for optimization of a simply

supported beam and a cantilever beam. In both test problems, the goal was to minimize the

compliance, or equivalently strain energy, subject to a volume constraint. The elastic modulus

Page 173: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

144

was set to 1 and the load applied was also 1; their magnitudes do not affect the optimal solution.

For the simply supported beam, also known as the MBB beam, a load is applied at the center of

the beam. The beam is standardized to a 6:1 ratio, or a 120 mm x 20 mm beam. In this validation,

and common to the structural optimization literature, half symmetry was used to reduce the

computational cost. For the ground structure, a 7x5 array of nodes represented the half symmetry

design space; see Figure 3.10. For the cantilever, the load is applied in the middle of the end; see

Figure 3.11. The design space is standardized to a ratio of 1.6:1 or a 40 mm x 25 mm beam. A

full 2D ground structure with 7x5 nodes represents the design space; see Figure 3.11.

Both benchmarks were solved using the minimization of strain energy objective and the

fmincon optimization algorithm in MATLAB. In the GSTO, it determines how to size each

element to minimize the compliance of the structure subject to the volume constraint. As the

solution progressed, the elements with the smallest cross-sectional areas approach the lower limit

and are later discarded from the final structure. For additional details on the benchmark problems;

see [273,336].

Page 174: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

145

Figure 3.10 Schematics for the MBB beam benchmark problem and the initial ground structure used

to represent the design space.

Figure 3.11 Schematics for the cantilever beam benchmark problem and the initial ground structure

used to represent the design space.

Page 175: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

146

For topology optimization, the solution is highly dependent on the original mesh and

constraints such as minimum feature size. For the ground structure model, the mesh density and

minimum feature size are comparable to the nodal array density and the relevant cross-sectional

area, respectively. The results of the GSTO of the MBB beam and cantilever benchmarks are

shown in Figure 3.12 and Figure 3.13, respectively. In the ALTO results, the line thickness in the

plot was correlated to the size of the cross-sectional area, thicker lines mean a thicker cross

section. Multiple solutions were optimized for a fixed allowable volume, while varying the

minimum allowable thickness. The minimum thickness value for the MBB beam and cantilever

benchmarks were (0.1, 0.5, 0.8) and (0.1, 0.5, 0.7), respectively. The compliance value of each

structure is noted on each image. In both test problems, the ground structure solutions were not

identical matches to those found in literature because the literature results are for a continuum-

based topology optimization as opposed to the discrete GSTO. However, they agree quite well

qualitatively; the trends in the topology of the minimum compliance structure are similar.

Figure 3.12 MBB beam comparison with results reported in literature. The different solutions are

shown for a fixed volume and changing the minimum feature size limit. The line width on the bottom

images is correlated to the cross-sectional area from the GSTO solutions. Top row of images adapted

from [273].

Page 176: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

147

Figure 3.13 Cantilever benchmark comparison with literature results. Various solutions are shown

for a fixed allowable volume and changing the minimum feature size limit. The line width on the

bottom images is correlated to the cross-sectional area from the GSTO solutions. Top row of images

adapted from [273].

The GSTO results follow similar patterns compared to results found in literature. As

shown, as the minimum feature size increases, the compliance value also increases, meaning a

less stiff structure. By adjusting the volume constraint as well as adjusting the minimum feature

size, the structure may be further refined. In addition, the ground structure model was a more

discrete representation of the design space than the continuum models. For example, the

cantilever example from literature had a smooth curved surface flowing from the supports to the

load which was difficult for the ground structure to replicate without increasing the node density.

By increasing the node density, the ground structure solutions could more closely approach the

literature results. The similarity between the GSTO and literature results is considered validation

of the ground structure formulation and optimization.

Page 177: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

148

3.2.4.2 Thermal Conductance Benchmark

The thermal conductance formulation and corresponding objective functions have been

developed for ALTO applications requiring specific thermal conductance properties from the

lattice structure. Through combination with other objective functions, it offers the opportunity for

unique multi-functional lattices. The thermal conductance formulation is validated through a

common topology optimization benchmark for a heat sink problem.

In the benchmark, a design space is uniformly heated with thermal insulation surrounding

the design space. At one point along the edge of the design space, a heat sink is applied. The

objective of the optimization is to maximize the heat transfer through the heat sink. A schematic

of the benchmark is shown in Figure 3.14 and more details are provided by Bendsoe and

Sigmund [281] and Lohan et al. [278]. Using ALTO, the benchmark was solved using a 2D

square (10 mm x 10 mm) design space with an array of 7x7 nodes; see Figure 3.14. The gradient-

based algorithm fmincon was used to solve the benchmark. To reduce the number of design

variables, a non-full ground structure was utilized by restricting the element length to 6 mm or

less.

Page 178: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

149

Figure 3.14 Schematic for thermal conductance benchmark from literature. The continuum topology

optimization approach (left) is shown compared to the GSTO approach (right).

The solution to the benchmark problem is shown in Figure 3.15. In addition, the solutions

from two other sources using a continuum approach for topology optimization are shown in

Figure 3.15. In the ALTO solution, the line thickness represents the relative thickness of each

strut. As shown, the solutions agree qualitatively, creating tree-like structures that branch out to

fill the design space. Common to all solutions in Figure 3.15 were the branches gradually tapering

and reducing in cross section as the distance from the heat sink increases. The branching and

tapering of the solution were characteristic of benchmark solutions reported in literature. As with

the previous benchmark, the nodal array, 𝐴𝑟𝑒𝑙 parameter value, and the maximum element size

can all be adjusted to make minor modifications to the solution. The similarities between the

benchmark results and literature results validate the thermal conductance formulation and

objective function for the optimization.

Page 179: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

150

Figure 3.15 Thermal conductance benchmark results from literature (left, middle) and from the

GSTO in ALTO show good correlation, validating the thermal conductance formulation and

optimization objective. Left and middle images adapted from [278] and [281], respectively.

3.2.4.3 Homogenization Benchmark

The final benchmark problem for validation of the ground structure model and

homogenization formulation was replication of results from the origin of the homogenization

approach [297,298]. In Sigmund’s work, he reported several 2D and 3D structures and their

constitutive matrices. For validation, these same structures were replicated in ALTO to compare

the calculated constitutive matrix against those reported by Sigmund.

Considering various examples from two sources by Sigmund [297,298], 6 different

example lattice structures were replicated as part of the validation. These structures include five

2D examples and one 3D example; see Table 3.3. In order replicate these structures in ALTO, the

same 2D square design space, discretized into a 4x4 nodal array was created. Similarly, a square

3D design space was discretized into a 4x4x4 nodal array for the 3D example problem. The exact

dimensions of the nodal array were not given and cross-sectional areas of all the elements were

not given in the source. The dimensions and assumed material properties of the elements were not

critical as the final result was normalized. However, these structures represent optimized

solutions and the exact cross-sectional areas of each individual truss element were not specified in

Page 180: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

151

the sources. As such, some variation was expected without the ability to exactly match all the

cross-sectional areas.

Table 3.3 Comparison of homogenization benchmark problems and ground structure calculation.

For these benchmark problems, the ground structure was an exact match to the reported constitutive

matrix, images adapted from [297,298]

Literature Benchmark

Reported Constitutive Matrix

Replicated Structure

Calculated Constitutive Matrix

[𝐸] = [1.00 0.00 0.000.00 0.00 0.000.00 0.00 0.00

]

[𝐸] = [1.00 1.00 0.001.00 1.00 0.000.00 0.00 0.00

]

[𝐸] = [1.00 1.00 0.001.00 1.00 0.000.00 0.00 0.00

]

[𝐸] = [1.00 1.00 0.001.00 1.00 0.000.00 0.00 0.00

]

[𝐸] = [ 1.00 −1.00 0.00−1.00 1.00 0.00 0.00 0.00 1.00

]

[𝐸] =

[ 1.0 1.0 1.0 0.0 0.0 0.01.0 1.0 1.0 0.0 0.0 0.01.0 1.0 1.0 0.0 0.0 0.00.0 0.0 0.0 0.0 0.0 0.00.0 0.0 0.0 0.0 0.0 0.00.0 0.0 0.0 0.0 0.0 0.0]

Table 3.3 shows the structure in the original source, the constitutive matrix, and the

ALTO structure, and the calculated constitutive matrix. In all cases, the constitutive matrices

were found to be a match to the results reported by Sigmund. For 2D examples, it was relatively

simple to determine the appropriate truss connectivity and appropriate element cross-sectional

areas. Only a single 3D example was completed due to the added complexity of determining the

appropriate connectivity and cross-sectional areas of each element. Using these six examples of

Page 181: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

152

homogenized materials, the ground structure model and homogenization formulation in ALTO

were validated.

3.3 Closing Remarks

This chapter details the basis and theory of ALTO, a systematic optimization approach

for unit cell topology optimization. As detailed in this chapter, there are a variety of objective,

constraint, and penalty functions that may be applied to tailor the optimization for specific

applications. Of particular importance is the contribution of this work toward the advancing field

of DfAM. Using various methods, AM optimization constraints and ground structure restrictions

can be directly incorporated into the optimization to improve manufacturability and reduce

interpretation steps to be able to print an optimized solution. Three examples of benchmark

problems have been demonstrated to validate various aspects of the ground structure formulation,

objective functions, and optimization. With the validation complete, the next chapters explore the

application of ALTO to several case studies including multi-functional unit cells, tailoring

compliance through homogenization, and improving powder removability.

Page 182: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

153

Chapter 4 Case Studies on Multi-functional Lattices

The focus of this chapter is demonstration of ALTO for multi-functional lattice

structures. The chapter includes a summary of the relevant background and motivation as well as

descriptions of the case studies performed. Two case studies are presented for demonstrating the

benefits of using ALTO to generate a novel lattice structures, specifically with the intent of

finding a balance between two conflicting objectives. The first case study is a 3D example with

the evolutionary optimization algorithm Borg. The second case study is a 2D example using the

gradient-based optimization algorithm fmincon. Each case study follows the same general flow of

presenting the motivation, a formal description of the optimization, and presentation of the

results. Afterward, there is a combined discussion of the results of the case studies and a few

closing remarks. The following objectives and sub tasks are addressed in this chapter:

3. Explore multi-functional lattices through case study on unit cell generation for thermal

conductance and minimum strain energy

3.1. Define constraints and process limitations

3.2. Generate optimal lattice structure with and without DfAM

3.3. Fabricate lattice structure with DfAM considerations

3.4. Evaluate solutions and revise ground structure topology optimization

4.1 Background and Motivation

Though the freedom of AM allows for complex and intricate structures to be

manufactured, the challenge remains to be able to effectively design components with such

structures in mind. In the author’s opinion, our ability to manufacture has surpassed our ability to

model and design to the fullest potential of AM. The inclusion of an appropriate lattice structure

in a component that enhances the macroscale performance becomes a metamaterial design

problem requiring novel systematic design and optimization tools such as ALTO.

Page 183: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

154

One method to approach the metamaterial design problem is through topology

optimization of the mesoscale structure. While there are numerous research groups exploring

methods for metamaterial design, the direct incorporation of AM constraints into the optimization

has not been addressed. As mentioned by Liu et al. [295], there is a need for further incorporation

of design for AM (DfAM) limitations in topology design. Plocher and Panesar [336] cited

integration of DfAM as a major bottleneck for a streamlined design process. In a review on

modeling approaches by Dong et al. [116], one of their primary findings was that lattice

structures were particularly promising with their potential to be multi-functional materials in

various engineering applications.

The purpose of these case studies is to explore ALTO for generation of multi-functional

lattice structures. Topology optimization and lattice structures have been of particular interest for

heat sinks and other heat transfer applications [118,183–185,278]. With AM, heat sinks can be

integrated into components and serve multiple purposes such as structural stiffness in addition to

including enhanced heat transfer topology. In these case studies, the focus is on using

optimization to balance two conflicting objectives: (1) minimization of strain energy and (2)

maximization of thermal conduction. The strain energy objective seeks to use all available

volume to maximize the stiffness of the structure against the loading configuration. The thermal

conduction objective seeks to use all available volume to maximize the thermal conductance of

the unit cell for the specified heat sink configuration. For additional details regarding the

objective functions, see Section 3.2.3.1.1 and Section 3.2.3.1.2. In these case studies, the loading

and heating configurations were specifically posed to be in conflict, such that the optimization

was forced to explore a balance between these two primary objectives.

Page 184: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

155

4.2 Case Study 1: Generation of a Novel 3D Multi-functional Lattice

4.2.1 Optimization Description

The purpose of Case Study 1 was a 3D example to demonstrate a multi-objective

optimization with an evolutionary algorithm and to show the effects of DfAM considerations. The

optimization algorithm for this example is Borg. A formal description of the optimization

formulation is given in Eqns. (4.1)-(4.5), with additional details available in Section 3.2.2.2. A

summary of the objectives, constraints, and penalties is shown in Table 4.1 and described in the

following paragraphs.

𝑀𝑖𝑛𝑖𝑚𝑖𝑧𝑒 (𝑊𝑆𝐸 ,𝑊𝑀𝑇𝐶 ,𝑊𝑉𝐹 ,𝑊𝑅𝐸)

𝑆𝑢𝑏𝑗𝑒𝑐𝑡 𝑡𝑜 𝐴𝑒 − 𝐴𝑙𝑜𝑤𝑒𝑟 ≥ 0 (4.1)

𝐴𝑒 − 𝐴𝑢𝑝𝑝𝑒𝑟 ≤ 0 (4.2)

𝐴𝑒 = 𝐴𝑙𝑜𝑤𝑒𝑟 𝑓𝑜𝑟 𝐴𝑒 < 𝐴𝑟𝑒𝑙 (4.3)

𝑊ℎ𝑒𝑟𝑒 𝑊𝑖 = 𝑚𝑖(𝑠𝑖𝑃 + 𝑤𝑖(𝑨, 𝑳)) (4.4)

𝑃 =∑𝑝𝑗(𝑨, 𝑳)

𝑛𝑝

𝑗

, 𝑓𝑜𝑟 𝑗 = 1…𝑛𝑝 (4.5)

Table 4.1 Case Study 1: 3D Multi-functional lattice optimization summary

Min. Strain Energy (𝑤𝑆𝐸) (𝑚𝑖, 𝑠𝑖) = (1,100)

Max. Thermal Conductance (𝑤𝑀𝑇𝐶) (𝑚𝑖, 𝑠𝑖) = (−1,100)

Min. Volume Fraction Error (𝑤𝑉𝐹) (𝑉𝑓∗, 𝑚𝑖, 𝑠𝑖) = (0.15,1,100)

Min. Relevant Elements (𝑤𝑅𝐸) (𝑚𝑖, 𝑠𝑖) = (1,100)

Symmetry YZ midplanes

𝑑𝑟𝑒𝑙 0.3mm

𝑑𝑙𝑜𝑤𝑒𝑟 , 𝑑𝑢𝑝𝑝𝑒𝑟 (9.5x10-5,1.5) mm

𝐿𝑙𝑜𝑤𝑒𝑟 , 𝐿𝑢𝑝𝑝𝑒𝑟 (2.1,15) mm

𝜃𝑙𝑜𝑤𝑒𝑟 , 𝜃𝑢𝑝𝑝𝑒𝑟 (0,35) degrees

DfAM Unsupported-S Penalty Optimized with and without

Borg Evaluations 100,000

Page 185: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

156

Objective 1 was to minimize the strain energy subject to the loading case; see Figure

4.1A and Eqn. (4.6). Objective 2 was to maximize thermal conductance out of the unit cell

through a single face-centered node; see Figure 4.1B and Eqn. (4.7). If each objective were

evaluated individually, one would expect elements to primarily be oriented vertically for

Objective 1 and horizontally for Objective 2. This example demonstrates finding a balance

between two objectives and is analogous to solving for a heat sink lattice with rigidity in one

direction. The case study considers the optimization of the unit cell individually and not global

heat transfer or stiffness across lattice structure patterned from this unit cell. Objective 3,

minimization of the volume fraction relative error, encourages a 15% volume fraction within the

unit cell; see Eqn. (4.8). Objective 4, minimization of relevant elements, was added to encourage

reduced unit cell complexity; see Eqn. (4.9).

wSE = 𝒖𝑇𝑲𝒖 (4.6)

𝑤𝑀𝑇𝐶 =𝑞𝑒𝑓𝑓

�̅� (4.7)

𝑤𝑉𝐹 =

|𝑉𝑓∗ −

𝑨𝑇𝑳𝑉𝑢𝑐

|

𝑉𝑓∗ (4.8)

𝑤𝑅𝐸 = 𝑐𝑜𝑢𝑛𝑡(𝐴𝑒 > 𝐴𝑟𝑒𝑙) (4.9)

Page 186: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

157

Figure 4.1 Case Study 1: 3D multi-functional lattice example problem, simultaneously optimizing for

(A) minimization of strain energy and (B) maximizing the thermal conductance

Beginning with an 8mm unit cell size, using a 5x5x5 cubic array of nodes, each node was

connected to every other node in the system using a truss element, for a total of 7750 elements or

design variables. Next, the optimization constraints were limits on the relevant diameter and

bounds on the upper and lower cross-sectional area. Ground structure restrictions were applied for

an element length range, build angle range, and symmetry across the Y and Z midplanes. The

element length restriction converted elements with a length range of 2.1-15 mm to void non-

design elements. The build angle restriction converted elements with a build angle larger than 0°

and less than 35° to void non-design elements. The symmetry restriction links symmetric

elements to a single design variable. Based on the ground structure restrictions, the total number

of design variables was reduced from 7750 to 96. For comparison of the effects of the DfAM

penalty, two identical optimizations were completed with and without application of the

unsupported‒S penalty.

Page 187: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

158

4.2.2 Optimization Results

The results of the case study show a solution with vertical members at the corners that

provides stiffness for the applied loading, while the horizontal members draw heat toward the

heat sink. Using a serial version of Borg, the computation times for the optimization with and

without DfAM were 5.5 hours and 5.4 hours, respectively. The heat sink structure resembles the

branching tree-like structure that reaches throughout the unit cell as described previously by the

benchmark problem; see Section 3.2.4.2. The ground structure restriction on element length

restricts the elements to those connecting only to its closest neighboring nodes. The resulting

solutions from the optimization, Design A (without DfAM) and Design B (with DfAM), are

shown in Figure 4.2. Design A is not manufacturable due to the unsupported horizontal elements

and would require extensive design interpretation and manipulation; see Figure 4.3. On the other

hand, by including the DfAM penalty, Design B shows a similar branching structure and is fully

self-supported.

Page 188: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

159

Figure 4.2 Solutions to the optimization problem from Case Study 1. Design A and Design B are the

solutions without and with the DfAM penalty, respectively.

Figure 4.3 Design A has several long horizontal overhangs reducing the manufacturability of the unit

cell. Incorporating a DfAM penalty into the optimization, the algorithm discovered Design B which is

fully self-supported. The images depict an XZ plane through the center of the unit cell topology for

improved visibility of some of the unsupported regions.

Page 189: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

160

Table 4.2 shows the objective function values (𝑊𝑖), for the two optimizations, without

DfAM (Design A) and with DfAM (Design B). The ideal result would be the lowest objective

values in each of these cases. Comparing the objective function scores, Design A and Design B

have equivalent strain energy and thermal conductance, despite one being infeasible for

manufacturing without support material. The difference in objectives in Table 4.2 comes in the

volume fraction and the number of relevant elements, where Design B is higher in each objective

than Design A to accommodate the added material/elements needed for a self-supporting unit

cell. In general, one would expect an improved strain energy score with added material. However,

the equivalent strain energy and thermal conductance score mean that the added support material

for Design B is used for its self-supporting features.

Table 4.2 Comparison of objective function values for two designs from Case Study 1. Design A is not

manufacturable due to long unsupported overhangs in the topology.

𝑾𝑺𝑬 𝑾𝑴𝑻𝑪 𝑾𝑽𝑭 𝑾𝑹𝑬 Self-Supported

Design A 90.54 -0.100 0.544 138 No

Design B 90.54 -0.100 0.570 168 Yes

In multi-objective algorithms, the optimization finds a Pareto-optimal set of solutions.

Because there are multiple objectives there are many solutions that can be considered optimal

because of the tradeoffs in the objective functions. The Pareto-optimal solutions are non-

dominated solutions, or solutions for which no other solutions are better in all objectives. As

described in Section 3.2.2.2, Borg utilizes 𝜖-box dominance. In this problem large 𝜖 values for

each objective were chosen (𝜖𝑆𝐸 , 𝜖𝑀𝑇𝐶 , 𝜖𝑉𝐹 , 𝜖𝑅𝐸) = (20,0.05,300,300) such that the algorithm

narrowed down the Pareto-optimal solutions and returned only a single design for each instance

(Design A and Design B); see [258] for additional details on 𝜖-box dominance. With lower

epsilon values than those specified for this problem, hundreds or even thousands of optimal

solutions may be returned by Borg and populate the Pareto front.

Page 190: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

161

For another instance of the optimization, the 𝜖 values were reduced to

(𝜖𝑆𝐸 , 𝜖𝑀𝑇𝐶 , 𝜖𝑉𝐹, 𝜖𝑅𝐸) = (0.05, 0.001,0.005, 0.5) to demonstrate optimization algorithm, resulting

in 84 different Pareto-optimal solutions. Figure 4.4 shows two graphs of the solution space for 84

Pareto-optimal designs: (A) modified thermal conductance versus strain energy and (B) modified

thermal conductance versus volume fraction. The optimization was searching for designs in the

lower left of each graph. In addition, the red point in each graph is the optimized solution, Design

B, from case study. For ease of viewing, the fourth objective is not depicted in these figures of the

solution space. As shown in the figure, a natural result of the 𝜖-box dominance is the stepped

solution space. Further reduction of the 𝜖 value for the modified thermal conductance objective

would reduce this stepped solution. From Figure 4.4A, the optimization has found a minimum

strain energy value (𝑊𝑆𝐸 ≅ 90.54), where the four corner pillars are at near the upper limit for

the cross-sectional area and there are multiple designs that can achieve this with varying levels of

thermal conductance. From Figure 4.4B, the optimization results show a clear tradeoff between

the volume fraction and thermal conductance.

Page 191: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

162

Figure 4.4 Graphs of the solutions space for an instance of the Case Study 1 Optimization. Each point

is a Pareto-optimal solution for (A) modified thermal conductance versus strain energy objectives

and (B) modified thermal conductance versus volume fraction objectives. The red point in each

graph represents Design B shown in Figure 4.2

One advantage of a multi-objective optimization is the opportunity for a designer to

explore the tradeoffs and select a design from the optimal solution set. To select a single design,

the Pareto optimal solutions have to be narrowed based on the intent of the design. For example,

the designer could desire a solution with 𝑊𝑀𝑇𝐶 ≤ −0.10, 𝑊𝑆𝐸 < 95 , and minimum 𝑊𝑉𝐹. Table

Page 192: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

163

4.3 shows the Pareto-optimal solutions from Figure 4.4 fitting this requirement and sorted by the

𝑊𝑉𝐹 objective, also included in this list is Design B from Figure 4.2. As shown, the design which

minimizes the volume fraction with the aforementioned requirements on the thermal conductance

and strain energy objectives is Design B.

Table 4.3 List of Pareto-optimal solutions with the modified thermal conductance objective less than

-0.10, strain energy objective less than 95, and the volume fraction objective sorted smallest to

largest. According to the requirements of the designer, the optimal solution would be Design B shown

in the first row in bold typeface.

𝑾𝑺𝑬 𝑾𝑴𝑻𝑪 𝑾𝑽𝑭 𝑾𝑹𝑬 Self-Supported

Top Design (Design B) 90.541 -0.10000 0.570345 168 Yes

90.629 -0.10002 0.745860 176 Yes

90.541 -0.10010 0.788925 168 Yes

94.458 -0.10008 0.849405 166 Yes

90.541 -0.10000 0.861188 166 Yes

90.541 -0.11048 0.991856 196 Yes

90.544 -0.11008 1.045903 180 Yes

91.109 -0.11071 1.096788 176 Yes

90.594 -0.11013 1.098463 178 Yes

90.541 -0.11009 1.111183 174 Yes

92.243 -0.12000 1.337346 198 Yes

91.077 -0.12043 1.393450 186 Yes

90.562 -0.12019 1.398234 188 Yes

90.562 -0.12010 1.416453 184 Yes

In Case Study 1, the overlap and self-supported penalties resulted in a unit cell topology

that was immediately manufacturable as opposed to another solution that would have required

extensive design interpretation. The optimized solution was formatted in a file type that is

readable using nTopology [235] to generate the full solid 3D geometric model. As demonstration

of the printability, the unit cell was printed with FFF using a single unit cell and then patterned

into a 2x2x2 array and fabricated using L-PBF. The L-PBF example was printed using a 2x scale

for visibility. Figure 4.5 shows each of the example prints that completed successfully without

any support material. The single unit cell was printed in PLA using standard processing

Page 193: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

164

conditions on an Afinia H480 3D Printer. To meet minimum feature size requirements of the FFF

process, the model was scaled up 5x so that a single unit cell was a 40 mm cube. The L-PBF

example was printed on a 3D Systems ProX DMP 320 using recycled Inconel 625 powder with

standard processing conditions.

Figure 4.5 Printed examples of the optimized solution from the 3D multi-functional lattice case study.

The examples were printed as a single unit cell (scaled up 5x) with FFF and in a 2x2x2 array (scaled

2x) with L-PBF. By incorporating DfAM considerations into the optimization, the optimized solution

was print-ready, not requiring any design interpretation.

4.3 Case Study 2: Generation of a Novel 2D Multi-functional Lattice

4.3.1 Optimization Description

Case Study 2 demonstrates a 2D example of the benefits of multi-objective optimization

with a gradient-based algorithm by solving the problem multiple times with different weighting

factors to show the tradeoffs between the objective functions. In this example, the two objective

Page 194: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

165

functions were the minimization of strain energy (𝑤𝑆𝐸̅̅̅̅ ) and maximization of thermal conductance

(𝑤𝑇𝐶̅̅̅̅ ). A formal description of the optimization problem is given in Eqns. (4.10)-(4.15),

𝑀𝑖𝑛𝑖𝑚𝑖𝑧𝑒 𝑓(𝑊)

𝑆𝑢𝑏𝑗𝑒𝑐𝑡 𝑡𝑜 𝐴𝑒 − 𝐴𝑙𝑜𝑤𝑒𝑟 ≥ 0 (4.10)

𝐴𝑒 − 𝐴𝑢𝑝𝑝𝑒𝑟 ≤ 0 (4.11)

𝑨𝑇𝑳 − 𝑉𝑓∗𝑉𝑚𝑎𝑥 ≤ 0

𝐴𝑒 = 𝐴𝑙𝑜𝑤𝑒𝑟 𝑓𝑜𝑟 𝐴𝑒 < 𝐴𝑟𝑒𝑙 (4.12)

𝑊ℎ𝑒𝑟𝑒 𝑓 = (𝑊) (4.13)

𝑊 =∑𝑐𝑖𝑚𝑖𝑤𝑖(𝑨, 𝒍), 𝑓𝑜𝑟 𝑖 = 𝑆𝐸̅̅̅̅ , 𝑇𝐶̅̅̅̅

𝑛

1

(4.14)

1 = ∑𝑐𝑖

𝑛

1

, 𝑓𝑜𝑟 𝑖 = 𝑆𝐸̅̅̅̅ , 𝑇𝐶̅̅̅̅ (4.15)

with additional details available in Section 3.2.2.1. A summary of the objectives, constraints, and

penalties is shown in Table 4.4 and described in the following paragraphs.

Table 4.4 Case Study 2: 2D Multi-functional lattice optimization summary

Min. Strain Energy (𝑤𝑆𝐸̅̅̅̅ ) (𝑚𝑖) = (1𝑥1010)

Max. Thermal Conductance (𝑤𝑇𝐶̅̅̅̅ ) (𝑚𝑖) = (−1)

Weight values for combined objective (𝑊) 𝑐𝑆𝐸 = 0,0.01,0.02,… ,1𝑐𝑇𝐶 = 1,0.99,0.98,… ,0

Volume Fraction Constraint 𝑉𝑓∗ = 0.1

Symmetry XY midplanes

𝑑𝑟𝑒𝑙 0.3mm

𝑑𝑙𝑜𝑤𝑒𝑟 , 𝑑𝑢𝑝𝑝𝑒𝑟 (1x10-6,2) mm

𝜃𝑙𝑜𝑤𝑒𝑟 , 𝜃𝑢𝑝𝑝𝑒𝑟 (1,35) degrees

Young’s Modulus E = 200 GPa

Thermal Conductivity 54 (W/m·K)

Coefficient of thermal expansion 1.2x10-5 (m/m·K)

Structural Load 4x: 1000N

𝑇1, 𝑇2 100°C, 25°C

For the minimization of strain energy, the load case is shown in Figure 4.6. Loads were

applied vertically downward on the four nodes at the top of the design space with pin supports in

Page 195: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

166

the bottom corners. The purpose of the first objective function, 𝑤𝑆𝐸̅̅̅̅ , was to minimize strain

energy due to loading in the negative Y-direction. While these conditions do not represent

periodic boundary conditions that would be present in a unit cell, the purpose was to illustrate the

minimization of strain energy of a lattice structure under the compressive loading. These loading

and boundary conditions may easily be varied by a designer depending on the application. The

purpose of the second objective function, 𝑤𝑇𝐶̅̅̅̅ , was to maximize thermal conductance in the X-

direction, from the left-edge nodes to the right-edge nodes; see Figure 4.6. A temperature

difference between the nodes on the left and right edges of the design domain was applied. The

unit cell optimization was solved using a 4x4 array of nodes for a square 5 mm design space and

assumes a ground structure restriction of symmetry across the vertical and horizontal midplane,

shown by the dashed lines in Figure 4.6.

Figure 4.6 Case Study 2: (A) Minimize strain energy and (B) Maximize thermal conductance. Blue

dashed lines denote symmetry planes. Images adapted from [361].

Because the objective function for the gradient-based optimization was a weighted sum

of the two objectives, a variety of weight values were required to find various solutions showing

the tradeoffs in the objectives. The weighting values (𝑐1 = 0,… ,1) and (𝑐2 = 1,… ,0), with

Page 196: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

167

increments of 0.01, were applied to each objective function, resulting in a set of 101 solutions. To

normalize the strain energy objective function to be of a similar order of magnitude as the thermal

conductance objective function, a scale factor of 1x1010 was applied to 𝑤𝑆𝐸̅̅̅̅ . The individual

objective functions and combined objective are shown in Eqns. (4.16)-(4.18). No penalty

functions were applied this optimization. DfAM considerations for the lattice structure

optimization were the ground structure restrictions on build angle and the relevant diameter limit,

shown in Table 4.4. The ground structure restrictions of build angle and symmetry reduced the

number of design variables from 120 to 30.

𝑤𝑆𝐸̅̅̅̅ =

𝑤𝑆𝐸𝑚𝑖𝑛 −𝑤𝑆𝐸

𝑤𝑆𝐸𝑚𝑖𝑛 −𝑤𝑆𝐸

𝑚𝑎𝑥 (4.16)

𝑤𝑇𝐶̅̅̅̅ =

𝑤𝑇𝐶𝑚𝑖𝑛 −𝑤𝑇𝐶

𝑤𝑇𝐶𝑚𝑖𝑛 −𝑤𝑇𝐶

𝑚𝑎𝑥 (4.17)

𝑊 = (1𝑥1010)𝑐𝑆𝐸𝑤𝑆𝐸̅̅̅̅ − 𝑐𝑇𝐶𝑤𝑇𝐶̅̅̅̅ (4.18)

4.3.2 Optimization Results

The result of the optimization was a set of 101 solutions showing the tradeoffs that exist

between the two objective functions. The total computation wall time was 2-3 hours using 6

computing cores. A convergence plot for the 𝑐𝑆𝐸 = 0.5 and 𝑐𝑇𝐶= 0.5 is shown in Figure 4.7.

Figure 4.8 illustrates 11 of the solutions with weight value increments of 0.1, showing the range

solutions possible. In each plot, the line widths of the elements were scaled based on its cross-

sectional area. Elements with a diameter less than 𝑑𝑟𝑒𝑙 were excluded from the plots. From Figure

4.8, it is shown that the optimal solution for maximizing the thermal conductance was comprised

of primarily horizontal bars while the optimal solution for minimization of strain energy was

comprised of primarily vertical bars. The result was that for a unit cell that has both high thermal

conductance and low strain energy, there must be a tradeoff.

Page 197: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

168

Figure 4.7 Example convergence plot for the solution to the optimization problem with equal

weighting factors for both objectives. The result shows an increasing objective function score until

the constraints are met and then gradually decreasing until convergence.

Figure 4.8: Case Study 2 results showing the tradeoffs between minimization of strain energy and

maximization of thermal conductance. 11 different solutions are shown beginning for incrementing

the weight values by 0.1 as the solutions move clockwise. Image adapted from [361].

Page 198: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

169

A graph of the objective function values for the 101 solutions generated is shown in

Figure 4.9. Optimal solutions would simultaneously maximize the thermal conductance and

minimize the strain energy. This figure shows the reciprocal of the thermal conductance objective

so the optimization searches for designs near the lower left corner of the plot. This figure

represents the designs obtained by incrementing the weighting factors and shows the Pareto front

that develops. The Pareto front is the set of designs for which no design is better than another in

both objectives, or non-dominated designs. These non-dominated solutions formed the Pareto

front and were all considered optimal. As one non-dominant design slightly improves in one

objective, it gets slightly worse in the other. Dominated designs are worse in both objectives than

at least one other solution. In Figure 4.9, non-dominated solutions are shown in grey circles and

dominated designs are shown as black ‘+’.

Figure 4.9: Comparison of the objective function values for minimization of strain energy and

maximization of thermal conductivity (minimization of its’ reciprocal plotted here). This plot shows

the solution space, demonstrating the tradeoffs between objectives. On the plot, all 101 solutions are

shown, with the grey dots being non-dominated solutions on the Pareto-front. Image adapted from

[361].

Page 199: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

170

4.4 Discussion of Results

The two case studies demonstrated in this chapter were for generating novel unit cell

topologies based on specific strain energy and thermal conductance targets. Though the case

studies were different, 3D versus 2D and Borg vs fmincon, both were successful in generating

novel unit cells and demonstrate unique advantages and disadvantages of each optimization

approach.

Through Case Study 1, ALTO demonstrates the ability to generate a manufacturable unit

cell topology for multiple conflicting design objectives. One challenge was the large number of

design variables in a full 3D ground structure. Even after reducing the optimization problem

through ground structure restrictions by considering symmetry and DfAM, the number of

variables was large, making it difficult for the algorithm to find a solution. One advantage of the

evolutionary algorithm Borg, was a true multi-objective optimization that can handle objectives

with different ranges and orders of magnitude. They can offer the additional flexibility in

returning a set of Pareto-optimal solutions allowing the designer to explore tradeoffs and select

from a set of optimal solutions. In addition, the result of the optimization clearly shows that

without inclusion of the DfAM penalty function, the unit cell topology would require additional

design interpretation and refinement. With the included DfAM penalty, we see tradeoffs arise in

the objectives due to the need for generating a self-supporting unit cell. The optimized solutions

could also be manufactured without any additional manipulation as shown by the printed

examples.

In Case Study 2, the 2D example illustrates the tradeoffs that exist between different

objectives because the weight values were given and the solution searches in a single direction.

One challenge of the 2D approach, using the gradient-based algorithm, was appropriately scaling

each objective function before the weighted sum was calculated for the final objective value.

Page 200: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

171

Initially, each objective function was normalized based on the range of the objective function

value; however, the minimization of strain energy objective has such a large range that the

normalized value was substantially smaller than the normalized thermal conductance value. If one

objective function is significantly smaller, or larger, than the other, then a single objective

function can become dominant and overshadow another objective. To rectify this difference,

significant time was required to determine an appropriate scaling factor for the minimization of

strain energy objective. With the scale factor applied, it placed each objective function value on a

similar order of magnitude however, this scale factor selection remains case-dependent and

somewhat arbitrary. This same issue exists for all optimization where multiple objectives and

penalties are combined into a single objective optimization through a weighted sum.

In general, there are several challenges and limitations of a GSTO approach, 1) the initial

nodal array restricts the design space to truss connected to those elements, 2) the assumption that

elements are connected with hinges of zero stiffness, and 3) finding a balance between the

number of elements, or design variables, and not overly restricting the design space.

In GSTO, the final solution is limited by the node and truss locations available in the

initial configuration. The nodal configurations were 5x5x5 and 4x4 for the 3D and 2D case

studies, respectively. These configurations were chosen based on recommendations from

literature. Sigmund [297] recommended that 4 or 5 nodes per edge were sufficient for a RVEs,

representing a good balance between not restricting the available design space and a reasonable

number of design variables. As future work, other ground structure approaches with moving

nodes could be considered to reduce the impact of the node array on limiting the design space and

compare the effect on the optimized solutions.

For the pin-joint connected truss elements, the stiffness at the nodes may be assumed to

be small when the structure carries the majority of the load through the elements themselves.

Deshpande et al. [201] found that hinge or pin joints can be a reasonable assumption. In addition,

Page 201: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

172

experimental results presented in Chapter 5 also demonstrate that the pin joint assumption was

reasonable. Where these assumptions break down is for thicker elements or high volume

fractions, where the size or prevalence of nodes within the unit cell increase relative to the void or

truss element space. For other approaches that account for nodal stiffness in unit cell design, the

reader may consider geometric projection methods [313–315].

Finally, there is a tradeoff between the complexity of the initial configuration and the

number of design variables included in the optimization. Apart from the nodal locations

themselves, the number of truss elements can become intractable. For a 5x5x5 nodal array there

are 7750 possible truss elements or design variables. While this allows for a wide range of

solutions, reducing 7750 design variables to a simple set of non-intersecting and manufacturable

elements was quickly intractable for multi-objective evolutionary algorithms. In this work,

ground structure restrictions such as symmetry and element build angle were used to reduce the

number of design elements. However, in comparison to continuum element approaches, 7750

elements are approximately equal to a cube discretized into 20 elements per edge, which would

not be capable of capturing the intricacy that can be accomplished through GSTO. As a result,

after simplification for symmetry and other restrictions, there was a significant reduction in the

number of design variables required compared to continuum topology optimization approaches.

As with all topology optimization methods, a final challenge is the interpretation step to

move from the optimized solution to the final design. This design interpretation is required for the

ground structure approach as with other continuum-based approaches. In the case of ALTO, Case

Study 1 was able to remove this interpretation step and generate a unit cell topology that was

manufacturable.

Page 202: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

173

4.5 Closing Remarks

In this work, ALTO has been demonstrated as an approach to improve unit cell DfAM,

including the ability to incorporate multiple design objectives into a single lattice structure. The

case studies shown in this work represent examples of metamaterial design for novel AM lattices

by incorporating stiffness in one direction, while transferring heat in an orthogonal direction.

Particularly with Case Study 1, the two solutions show the importance of including DfAM

considerations into the optimization. Without direct integration of DfAM constraints, penalties or

ground structure restrictions, design interpretation to ensure manufacturability would be essential.

For Case Study 2, the wide range of solutions along the Pareto-front demonstrate the importance

of including multiple objectives that can allow for the tradeoffs in designs. Through continued

improvement, ALTO offers an approach to take advantage of the intricate design freedom offered

through AM, including the development of multi-functional lattices or metamaterials.

Page 203: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

174

Chapter 5 Case Studies on Homogenized Lattice Structures

The focus of this chapter is exploration of ALTO for generating unit cells with tailored

elastic properties. The chapter includes a brief summary of the relevant background and

motivation as well as descriptions of the case studies performed. Three case studies are presented

for demonstrating the benefits of using ALTO to generate a novel lattice structures, specifically

with the intent of matching a target homogenized constitutive matrix. The first case study is a 3D

example exhibiting improved powder removability. The second case study is a 3D example for

creating a unit cells with different stiffness by changing topology without decreasing the volume

fraction. The third case study is a 2D example for creating an orthotropic material that is three

times stiffer in one direction than the other. Each case study follows the same general flow of

presenting the motivation, a formal description of the optimization, and presentation of the

results. Afterward, there is a combined discussion of the results of the case studies and a few

closing remarks. The following objectives and sub tasks are addressed in this chapter:

4. Explore homogenized lattices through case study on generation of unit cells with tailored

compliance

4.1. Define constraints and process limitations

4.2. Generate optimal lattice structure with tailored properties

4.3. Fabricate and test to determine mechanical properties

4.4. Characterize fabricated samples with mechanical testing

4.4.1. Compare boundary conditions of homogenization with experimental

characterization

5.1 Background and Motivation

Lattice structures present a unique opportunity for materials with tailored

properties/behavior. Topology design of cellular materials has been ongoing for over thirty years,

dating back to the significant contributions of Bendsoe and Sigmund [268,298]. In these works,

Page 204: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

175

the mesostructure or cellular structure of a component was designed to match specific elastic

constitutive properties through homogenization. Through their work, they were able to

demonstrate mesostructures with extreme mechanical properties, such as designs with Poisson’s

ratios in the range (-1,1) [297]. At the time of this early research, the limited manufacturing

options were recognized as a primary challenge and rapid prototyping was mentioned as a

potential prospect. With recent advances in AM, there is renewed interest in design of cellular

structures, or lattice structures.

For L-PBF, there are several challenges with printing lattice structures which are

summarized here; for additional details, see Section 1.1.4.2 and Section 1.1.5.1. In their review of

lattice structures fabricated with AM, Nazir et al. [244] identified four challenges: (1) support

material removal, (2) powder removal, (3) large file size and (4) high surface roughness. In

addition to the four limitations covered by Nazir et al. [244], Maconachie et al. [250] address the

staircase effect which can cause significant changes in cross section for thin lattice features and

describe the impact of these challenges on mechanical properties. The ability to design unique

lattice structures with tailored properties, while accommodating DfAM is a metamaterial problem

requiring optimization.

The purpose of this chapter is demonstration of ALTO for generation of unit cells that

have a specific target constitutive matrix. The unit cell topology for a lattice structure is

frequently chosen from a list of common structures, without understanding of how the unit cell

topology will affect the overall properties of the component. In these case studies, the desired

lattice structure properties were assumed to be known as a target constitutive matrix and then

ALTO was used to generate novel unit cell topologies that matched that constitutive matrix.

Three unique case studies are presented to demonstrate the abilities of ALTO. Case Study 1 is a

3D example highlighting a novel powder removability objective. The optimized solution matches

the orthogonal effective elastic moduli of a common lattice structure but exhibits improved

Page 205: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

176

powder removability. Case Study 2 demonstrates the generation of three novel 3D unit cell

topologies by scaling the constitutive matrix for fixed volume fraction. Case Study 3

demonstrates generation of an orthotropic unit cell topology that is three times stiffer in one

direction than the other. Using a multi-objective optimization approach, such as ALTO, to design

unit cell topologies with a target constitutive matrix enables metamaterial design with enhanced

functionality.

5.2 Case Study 1: Generation of a 3D Unit Cell Topology with Improved Powder

Removability

The purpose of Case Study 1 was to demonstrate ALTO for generation of a new unit cell

that has the same orthogonal effective moduli as a baseline unit cell and improved powder

removal. After optimization, the optimized and baseline unit cell topologies were printed and the

mechanical properties were compared through compression testing. Powder removal was

evaluated through comparison of measured weight of the as-designed and as-printed components

and further validated through CT-scan evaluation.

The baseline unit cell chosen for the study was an octet-truss, shown in Figure 5.1. The

octet-truss is an arrangement of 36 struts that create a large number of potential trapped powder

regions. For this study, the unit cell size was 8 mm with 1 mm struts, or a volume fraction of

17.1% based on the CAD model (20.8% in ALTO). The octet-truss unit cell [362], chosen as the

baseline unit cell, is very common in literature because the effective modulus was predicted to

scale linearly with volume fraction [141,201]. This linear scaling was recently demonstrated in

[363]. Other studies have also used the octet-truss as a baseline for comparison against other unit

cells that follow similar linear scaling trends [133,301,363]. Mechanical properties that have been

studied include static effective modulus [150,191,364–366], high-strain rate compression

Page 206: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

177

[191,367], and fracture toughness [157]. Various other groups have studied it for optimization

[368], medical cellular scaffolds [176], or nanoscale applications [369].

5.2.1 Optimization Description

For the optimization, four objectives were included: 1) minimize homogenization error,

2) minimization of volume fraction, 3) minimization of the relevant elements in the unit cell, and

4) minimize the powder removability factor (PRF). The optimization constraints were upper and

lower bounds on the cross-sectional area and the relevant diameter limit for the minimum feature

size. Ground structure restrictions were the build angle limit, symmetry across ‘X’, ‘Y’, and ‘Z’

midplanes of the unit cell, and interconnectivity nodes. No penalties were applied for this

optimization. The optimization algorithm was Borg with a stopping criterion of 70,000 unit cell

design evaluations. A formal description of the optimization is shown in Eqns. (4.1)-(5.4). A

summary of the optimization problem is shown in Table 5.1. The optimization objectives and

constraints are described briefly below. For additional details, see Chapter 3.

𝑀𝑖𝑛𝑖𝑚𝑖𝑧𝑒 (𝑊𝐻 ,𝑊𝑉𝐹 ,𝑊𝑅𝐸 ,𝑊𝑃𝑅𝐹)

𝑆𝑢𝑏𝑗𝑒𝑐𝑡 𝑡𝑜 𝐴𝑒 − 𝐴𝑙𝑜𝑤𝑒𝑟 ≥ 0 (5.1)

𝐴𝑒 − 𝐴𝑢𝑝𝑝𝑒𝑟 ≤ 0 (5.2)

𝐴𝑒 = 𝐴𝑙𝑜𝑤𝑒𝑟 𝑓𝑜𝑟 𝐴𝑒 < 𝐴𝑟𝑒𝑙 (5.3)

𝑊ℎ𝑒𝑟𝑒 𝑊𝑖 = 𝑚𝑖(𝑤𝑖(𝑨, 𝑳)) (5.4)

Page 207: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

178

Table 5.1 Case Study 1: Improved powder removal lattice optimization summary

Min. Homogenization Error (𝑚𝑖) = (1)

Min. Volume Fraction Error (𝑉𝑓∗, 𝑚𝑖) = (0.01,1)

Min. Relevant Elements (𝑚𝑖) = (1)

Min. PRF (𝑝∗, 𝑚𝑖) = (4𝑚𝑚, 1)

Symmetry XYZ midplanes

𝑑𝑟𝑒𝑙 0.3 mm

𝑑𝑙𝑜𝑤𝑒𝑟, 𝑑𝑢𝑝𝑝𝑒𝑟 (9.5x10-5,3) mm

𝜃𝑙𝑜𝑤𝑒𝑟, 𝜃𝑢𝑝𝑝𝑒𝑟 (1,30) degrees

Borg Evaluations 70,000

The first objective used homogenization methods [116,297,298] to predict the

constitutive matrix and minimize the error between the calculated and target constitutive matrix.

The target constitutive matrix, based on the octet-truss baseline cell (volume fraction of 20.8%),

is shown in Eqn. (5.5), where �̃�∗ is the target constitutive matrix, and E is the material Young’s

modulus. The target constitutive matrix differs from octet-truss homogenized constitutive matrix

by a 25% reduction in each of the three orthogonal shear terms (𝐸44𝐻 , 𝐸55

𝐻 , 𝐸66𝐻 ). The objective

function is shown in Eqn. (5.6), where 𝑤𝐻 is the raw objective score and �̃�𝐻 is the calculated

homogenized constitutive matrix unit cell design.

�̃�∗ = 𝐸

[ 𝟎. 𝟎𝟑𝟒𝟕 𝟎. 𝟎𝟏𝟕𝟒 𝟎. 𝟎𝟏𝟕𝟒 𝟎 𝟎 𝟎

𝟎. 𝟎𝟑𝟒𝟕 𝟎. 𝟎𝟏𝟕𝟒 𝟎 𝟎 𝟎𝟎. 𝟎𝟑𝟒𝟕 𝟎 𝟎 𝟎

𝟎. 𝟎𝟏𝟕𝟒 𝟎 𝟎𝟎. 𝟎𝟏𝟕𝟒 𝟎

𝒔𝒚𝒎 𝟎. 𝟎𝟏𝟕𝟒]

(5.5)

𝑤𝐻 = ∑

|(𝐸𝑘𝑙∗ −𝐸𝑘𝑙

𝐻)|

𝐸𝑘𝑙∗ for 𝑘, 𝑙 = 1…6 (5.6)

The second objective minimizes the error between the target and current volume fraction.

Eqn. (5.7) shows the volume fraction objective function 𝑤𝑉𝐹, where 𝑨 is a vector of the element

cross-sectional areas, 𝑳 is vector of element lengths, and 𝑉𝑓∗ is the target volume fraction. Note

that the volume fraction calculated in ALTO overpredicts the actual volume fraction because it

Page 208: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

179

does not account for overlapping volumes near nodes. In this case study, the volume fraction was

set to 1% to effectively minimize the volume fraction. The third objective encourages simpler

unit cell designs with fewer struts; see Eqn. (5.8), where 𝐴𝑒 is the element cross-sectional area

and 𝐴𝑟𝑒𝑙 is the relevant cross-sectional area limit or minimum feature size. Though reducing

complexity is not often a focus for AM, at the mesoscale, the strut diameters and proximity of

struts within a unit cell are so close to the minimum feature size of industrial machines that

reduced complexity can improve print quality in lattice structures.

𝑤𝑉𝐹 =

|𝑉𝑓∗ − 𝑨𝑇𝑳|

𝑉𝑓∗ (5.7)

𝑤𝑅𝐸 = 𝑐𝑜𝑢𝑛𝑡(𝐴𝑒 > 𝐴𝑟𝑒𝑙) (5.8)

The fourth objective was to minimize the PRF. For each unit cell design, the pore

distance matrix, �̃�, was assembled based on the midpoint to midpoint distance between all

elements in the initial ground structure, �̃�𝒆𝒆, minus the radius of each element, 𝑟𝑖, 𝑟𝑗 ; see Eqn.

(5.9). After assembling �̃�, the objective function was calculated by counting the number of

instances that fell below the pore threshold, 𝑝∗; see Eqn. (5.10). For the fourth objective, the

threshold value for the pore size was set to 4 mm, or 50% of the unit cell size, which is much

larger than the powder particle size. For a large unit cell size and low volume fraction, such as the

baseline octet-truss, the purpose of the objective is to minimize the number of pores smaller than

4 mm in an 8 mm unit cell for improved powder flowability out of a lattice. For a small unit cell

size or high volume fraction of the optimized topology, the 𝑝∗ value will also help reduce

potential trapped regions.

𝑃𝑖𝑗 = 𝐷𝑖𝑗

𝑒𝑒 − 𝑟𝑖𝑒 − 𝑟𝑗

𝑒 , 𝑓𝑜𝑟 {𝑖 = 1,… ,𝑀𝑗 = 1,… ,𝑀

} (5.9)

𝑤𝑃𝑅𝐹 = 𝑐𝑜𝑢𝑛𝑡(𝑃𝑖𝑗 < 𝑝∗) (5.10)

Page 209: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

180

For optimization constraints, the upper and lower bounds on the cross-sectional area and

the relevant cross-sectional area limit were based on the equivalent diameter for the circular cross

section as defined in Table 5.1. Ground structure restrictions included build angle limit,

interconnectivity nodes, and symmetry. The unit cell was modeled by a full ground structure for a

5x5x5 cubic node array. Next, with the exception of 5 nodes per face (four corners and the center

of each face), all nodes on the surface of the unit cell and their corresponding elements were

converted to void non-design elements which have a fixed cross-sectional area at the lower limit.

This ground structure restriction mimics the nodes on the surface of the octet-truss unit cell; see

Figure 5.1. In addition, based on the ground structure restriction for build angle limit, any

elements between a build angle of 1 and 30 degrees (defined from the build plate) were converted

to void non-design elements. For ALTO, horizontal elements that form bridges between nodes

were allowed because the short distance they spanned across the unit cell is generally printable.

Finally, the symmetry ground structure restriction across all three orthogonal midplanes was

applied for cubic symmetry in the unit cell, linking all symmetric element cross-sectional areas to

a single design variable in the optimization. Beginning with a total of 7750 design variables, the

ground structure restrictions on interconnectivity nodes, build angle limit, and symmetry reduced

the total number of design variables for the ground structure to 120.

Page 210: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

181

Figure 5.1 The octet-truss (A) is the baseline unit cell topology for matching the constitutive matrix,

colors have been added to improve visualization of 3D topology. The initial ground structure is

shown in (B), where each line represents a truss element. With symmetry, there were a total of 120

distinct design variables.

5.2.2 Optimization Results

The optimization problem was successful in identifying multiple unit cell topologies with

𝑤𝑆𝐸 < 10%, or less than 10% total relative error between the calculated and target constitutive

matrix. From the designs with less than 10% total relative error, two unit cell topologies were

identified based on their ability to exactly match the target constitutive matrix. The two unit cells

from the optimization were the octahedral and a new cell type called the opti-octet. Figure 5.2

shows a single unit cell and a side view of the patterned structure for the two unit cells from the

Page 211: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

182

optimization on either side of the baseline octet-truss unit cell. A summary of the objective

function values is shown in Table 5.2.

Figure 5.2 The two unit cells output from the optimization are shown on either side of the octet-truss.

The top images show a single unit cell and the bottom images are side views of a 3x3 patterned array.

As a result of the PRF as an objective, both the octahedral and opti-octet open up the porous space to

improve powder removal and flowability.

Table 5.2 Summary of the objective function values for the optimization results. The octet-truss

results are shown for individual evaluation of the unit cell, separate from the optimization.

Objectives Opti-Octet Octet-Truss Octahedral

𝑤𝐻 0% -- 0%

𝑤𝑉𝐹 20.8% 20.8% 20.8%

𝑤𝑅𝐸 20 36 12

𝑤𝑃𝑅𝐹 36 252 60

For the two unit cells identified from the optimization results, minor adjustments to the

optimized solutions were made as part of the design interpretation step. One challenge with non-

Page 212: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

183

gradient-based optimization algorithms is that they lack an intelligence about the model. Rather

than progressively increasing/decreasing a design variable value based on a gradient

approximation, non-gradient-based algorithms rely on various recombination methods. These

recombination methods can make it difficult for minor variations of an optimal solution when

recombining with other designs. In the optimization results, the octahedral and opti-octet had 𝑤𝑆𝐸

values of 6.65% and 2.31%, respectively. In each case, manual adjustments to the strut diameters

were found to result in an exact match for the constitutive matrix and volume fraction as

compared to the baseline unit cell topology. The average manual adjustment to strut diameters

was 0.7% and 2.3% for the octahedral and opti-octet, respectively. This interpretation step is

common to topology optimization. The octet-truss baseline unit cell has a 75% difference from

the target constitutive matrix because of the reduction of the primary shear terms (𝐸44𝐻 , 𝐸55

𝐻 , 𝐸66𝐻 )

by 25% each, but matches exactly in all other terms of the constitutive matrix.

For multi-objective optimization, there are many designs which may be considered

optimal or said to lie on the Pareto front, where neither design dominates another in all objectives.

For example, the octahedral has fewer relevant elements but a larger number of potential trapped

regions than the opti-octet. On the other hand, the opti-octet has more relevant elements but fewer

potential trapped regions than the octahedral. With these tradeoffs in the objectives, these were

two solutions along the Pareto front.

From the side view in Figure 5.2, one can see that both the opti-octet and octahedral open

up larger pathways for powder flowability than the baseline octet-truss. The octahedral primarily

accomplishes large channels for powder flowability as a result of the minimization of relevant

elements objective. Having few elements in the unit cell structure creates wide open channels

with neighboring unit cells. However, the tight cavity at the center of the unit cell leads to the

high PRF score. On the other hand, the opti-octet accomplishes large channels for powder

flowability by spreading out the elements within the unit cell as shown by PRF. Based on the

Page 213: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

184

primary intention of improved powder removability, the opti-octet was chosen for fabrication and

further analysis in comparison with the octet-truss.

5.2.3 Ground Structure Model Validation

5.2.3.1 Fabrication and Testing

For validation of the ground structure model used in ALTO, a set of octet-truss and opti-

octet lattices were printed and evaluated using uniaxial compression testing. For both structures,

the unit cell size was scaled down to 4 mm and arrayed into a 10x10x10 lattice structure or 40

mm x 40 mm cubic samples. The intent of the 10x10x10 lattice was to reduce boundary effects

from the compression platens on the top and bottom of the sample as recommended in ISO 13314

[370]. The completed build plate is shown in Figure 5.3 with all the octet-truss and opti-octet

lattices among two other lattice types. The location of each lattice on the build plate was

randomized to minimize any effects from build location. The strut diameters for the octet-truss

were a uniform thickness of 0.5 mm. For the opti-octet, it is a unique combination of a cubic and

body-centered cubic (BCC) lattice, with strut diameters of 0.595 mm and 0.678 mm, respectively.

For testing, four replicates of each lattice were printed on an EOS M290 in Ti64 using standard

laser process parameters with infill only, no contours. The parts were heat treated at 800 C for

two hours in an argon inert atmosphere as recommended per the EOS material datasheet. The

parts were removed from the build plate using wire electrical discharge machining.

Page 214: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

185

Figure 5.3 Complete build plate with octet-truss and opti-octet lattice structures among a few other

lattice types. In the picture, one of the five samples of each lattice type are labeled. The interlocked

and diamond lattice are discussed as part of Case Study 2 in Section 5.3.

Using an MTS QTest/100 test frame, each sample was loaded in compression until the

first peak force at a crosshead speed of 1.0 mm/min, an equivalent strain rate of 4.17x10-4[1/s].

Figure 5.4 shows an example of the compressed lattice structures. From the load displacement

data, the global stress was calculated as the force divided by the original cross section of the

sample (1600 mm2). The global strain was calculated based on the crosshead displacement with

compensation for the compliance of the test frame.

Page 215: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

186

Figure 5.4 Examples of the octet-truss (left) and opti-octet (right) crushed after compression loading.

Prior to compression, the samples were 40 x 40 mm cubes or a unit cell size of 5mm as noted by the

scale bars.

5.2.3.2 Experimental Results

The average stress-strain curves from four samples each for the octet-truss and opti-octet

lattice structures are shown in Figure 5.5. The grey band behind each curve in the figure

represents the average plus and minus the standard deviation of the experimental data. The

effective modulus of each was computed by a linear least squares fit to the data between 10-50%

of the peak load. In Figure 5.5, the effective modulus region for each lattice is denoted by the

solid portion overlaid on the stress strain curves. From this region, the effective modulus was

calculated to be 2420.6 ± 65.4 MPa and 2152.4 ± 73.2 MPa for the octet-truss and opti-octet

lattices, respectively. The peak stress was determined to be 50.7 ± 1.5 MPa and 40.8 ± 0.9 MPa

for the octet-truss and opti-octet lattices, respectively. Table 5.3 shows a comparison of the

experimental, ALTO, and ALTO+corrections results of the effective modulus.

Page 216: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

187

Figure 5.5 Average stress strain curves for the octet-truss and opti-octet truss. The solid portion of

the curves represents the region where the effective modulus was measured, 10-50% of the peak

stress. The grey band behind each curve represents the average plus and minus the standard

deviation of the experimental data.

Table 5.3 Experimental validation results for the effective modulus and comparison to the ground

structure model prediction in ALTO and AM-based corrections. For AM corrections, IO = infill-only

correction and SC = stair case correction. The error relative to the experimental value is noted in the

parenthesis.

Octet-truss (MPa) Opti-octet (MPa)

Experimental 2420.6 ± 65.4 2152.4 ± 73.2

ALTO Prediction 2661.1 (+9.9 %)

2660.4 (+23.6 %)

ALTO + BOC 2279.4 (-5.9 %)

2276.4 (+5.8 %)

ALTO + BOC + SCC 2133.1 (-11.9 %)

2260.1 (+5.0 %)

From the experimental data, the ground structure model was found to slightly overpredict

the effective modulus and was attributed to differences between the as-designed and as-printed

models. There were several assumptions in the ground structure model and variations in the as-

printed part that can explain the discrepancy. For the ground structure, the model assumed perfect

geometry without deviations due to surface roughness or the scan strategy, that all nodes were

Page 217: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

188

representative of pin-joints, and that the stiffness of a unit cell was solely attributed to axial

loading of the printed struts. Deshpande et al. [201] showed that the pin-jointed truss element

assumption was a valid assumption, meaning that the axial stiffness of a strut plays a much

greater role in the effective modulus than the joint or bending stiffness at the nodes. In addition,

previous studies have shown that the number of layers of lattice changes the effective modulus of

the lattice due to the constrained, non-sliding, boundary conditions created by friction at the

surface of the compression platens [207]. For the results presented in Table 5.3, the experimental

data is compared against a homogenization prediction using “relaxed” boundary conditions that

represent the least constrained layers of the lattice sample, centrally located between compression

the platens. Additional detail on a comparison of the boundary conditions is given in Section

5.2.3.3.

For printing variations, the AM process is susceptible to deviations in geometry due to

the STL file conversion, the stair stepping effect, variable surface roughness depending on the

build angle, and the laser scan strategy or parameters to recreate the intended geometry. It is

assumed that powder removal does not affect the failure mechanisms in the lattice structures.

Two AM-based model corrections were considered for implementation in the ground structure: 1)

the Beam Offset Correction (BOC) and 2) the Stair Case Correction (SCC). The BOC is an offset

correction based on the laser processing parameters for the infill of each region to be scanned

with the laser in a layer. Lattice structures have a large surface to volume ratio, meaning that the

build time is a more evenly split between laser scan time for infill versus laser scan time for

contours as compared to a non-lattice part. In addition, scanning both infill and contours tends to

cause overheating in the thin lattice struts. In this case, the build time was decreased by nearly

50% by using an infill-only scan strategy as compared to including both the infill and contour

laser scans. As a result of using infill-only laser scans to build the lattice, the laser processing

parameters for Ti64 on an EOS M290 introduced a 0.025 mm beam offset from the as-designed

Page 218: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

189

cross-section in the XY build plane due to the laser scan strategy. The beam offset in the XY

plane caused a reduction in the cross-section which was dependent on the build angle of the strut.

Similarly, the BOC reduces the cross-sectional area of each strut in the lattice to match the actual

geometry scanned by the laser during the build process. Figure 5.6 shows a schematic of the

reduction caused by the beam offset and the adjusted cross-sectional area calculation.

The SCC is an additional offset correction to account for the staircase effect that is

frequently described for the high surface roughness on angled features. For lattice structures, it

has been suggested that the staircase effect can have a large impact on the stiffness of the lattice

struts because the roughness is large relative to the diameter of the struts [194,223,250,365].

Figure 5.6 shows how the staircase effect can negatively impact the cross-sectional area, by

reducing the continuous cross-sectional area over which a load is transferred. In the illustration,

underneath the overhang from each layer shift, there is dross buildup but the effectiveness of the

dross is limited, causing the reduction in cross section. Though the staircase effect has been

mentioned in several studies, a build angle dependent formula for approximating the effective

cross-sectional area of struts has not been used as a correction factor. The SCC was only applied

to angled struts, meaning struts with a non-vertical and non-horizontal build angle. The geometry

change for these two corrections was applied first as just the BOC, based on scan parameters, and

second, BOC combined with the SCC.

Page 219: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

190

Figure 5.6 AM-based corrections applied to ground structure prediction. As a result of either or both

corrections, the cross-section of the strut is effectively reduced.

The variation in percent error relative to the experimental values stemmed from the

nature of the unit cell types. First, consider the impact of the AM corrections based on strut angle.

The BOC correction affects all struts, with the biggest impact on vertical struts. On the other

hand, the stair case effect impacts only angled struts, with no impact on vertical or horizontal

struts. The BOC correction was first applied to account for the beam offset for all cross-sections

in the build plane. Next, the SCC correction was applied to adjust the effective cross section

based on the stair case effect of AM. For the opti-octet, it was apparent that the vertical and

horizontal struts carry the majority of the load because the BOC had a large impact on the

effective modulus while the SCC had very little impact on the effective modulus. On the other

hand, for the octet-truss, BOC decreased the effective modulus and underpredicts the

experimental result and then the added SCC reduces the effective modulus even further.

It is difficult to generalize an AM correction factor to account for multiple unit cell

topologies because there are many factors at play in the assumptions of the ground structure and

the AM geometric variations are geometry dependent. However, comparing the total error from

Page 220: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

191

Table 5.3, or combined error for both unit cell types, the BOC was found to be a sufficient

correction. While the SCC applied to the opti-octet effective modulus prediction was shown to

slightly reduce the relative error, in other cases, such as the octet-truss, the correction may lead to

an overcorrection. As such, the SCC was found to be an unnecessary correction for ground

structure models. In summary, the ground structure model, combined with the BOC was shown to

be a validated model that accurately predicted the effective modulus of two unique unit cells to

within 6%.

5.2.3.3 Comparison of Boundary Conditions

During compression, lattice layers near the upper or lower compression platens will

experience greater constraint than the centrally located layers due to the edge effects from friction

between the sample and platens. Figure 5.7 shows a schematic of a compression sample, with the

color gradient representative of the impact of edge effects and two examples of representative

boundary conditions for the unit cells in the “constrained” or “relaxed” regions of the

compression sample. Because globally the experimental testing measures the least stiff region, or

“relaxed” middle layers of the sample, the homogenization should predict the stiffness at the

central or “relaxed” region of the lattice.

Page 221: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

192

Figure 5.7 Schematic of a compression sample where the color gradient represents the impact of edge

effects from the friction with compression platens. Near the compression platens a unit cell

experiences “constrained” boundary conditions as compared to “relaxed” boundary conditions for a

unit cell near the middle of the sample.

Based on the homogenization method from Sigmund [297,298] the original boundary

conditions that were established in the homogenization objective function for the optimization

were found to be highly constrained over the experimental results. In the original boundary

conditions, the unit cell was constrained such that during the test strain cases, edges and faces

cannot contract as would be expected due to a Poisson’s ratio-like effect. From Section 3.2.3.1.3,

Figure 3.3 shows the boundary conditions for the 2D test strain cases. Similarly, in 3D the

boundary conditions restrict motion of the unit cell during the test strain cases such that the

effective modulus of the homogenized constitutive matrix may be overpredicted. For compression

testing a single layer of lattices, the unit cells would be highly constrained and expected to more

closely match the original boundary conditions as described in Section 3.2.3.1.3. Testing with a

minimum of 10 lattice layers, as suggested by ISO 13314 [370], produce a “relaxed” boundary

condition near the center of the sample. By relaxing the boundary conditions to mimic those

shown for the “relaxed” unit cell from Figure 5.7, the homogenized properties are significantly

Page 222: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

193

reduced. Table 5.4 shows the comparison of the effective modulus calculated from experimental

data and for two different boundary conditions in ALTO.

Table 5.4 Effective modulus comparison between experimental data and ALTO predictions of

“constrained” and “relaxed” boundary conditions. The error relative to the experimental value is

noted in the parenthesis.

Octet-truss (MPa) Opti-octet (MPa)

ALTO – “Constrained” Boundary Conditions

3991.7 (+64.9 %)

3991.8 (+85.5 %)

ALTO – “Relaxed” Boundary Conditions

2661.1 (+9.9 %)

2660.4 (+23.6 %)

Experimental 2420.6 ± 65.4 2152.4 ± 73.2

For matching the experimental data in Table 5.3 and validating the ground structure

model, ALTO predictions of the effective modulus were calculated using “relaxed” boundary

conditions that allow for expansion of the sides of the unit cell, similar to a Poisson’s ratio effect

in a traditional material. Using “relaxed” boundary conditions, and applying AM corrections

ultimately provides a very accurate prediction of the effective modulus of the compression

samples. Additional discussion of the issues that arise due to the choice of boundary conditions

for homogenization is given in Section 5.5.

5.2.4 Powder Removability

For comparison of powder removal between the octet-truss and the opti-octet, samples

fabricated with AM were weighed to compare the as-designed versus as-printed weight. The

ground structure model was used to compare the effective modulus prediction of the two unit

cells. If all powder was removed from the structure, the relative error between as-designed and

as-printed weight should be small and increase as trapped powder or partially sintered powder

exists inside the lattice structure. The purpose of the experiment was to demonstrate that the PRF

Page 223: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

194

objective function generated a unit cell with the same orthogonal effective moduli properties as

the octet-truss unit cell topology, but with improved powder removability.

5.2.4.1 Fabrication of Powder Removal Samples

For each unit cell type, 12 different test specimens were created by changing the unit cell

size while holding the volume fraction constant. This provided a set of 12 samples where each

sample was predicted to have the same effective modulus as its paired size in the other unit cell

type and the spacing between struts where powder can escape gradually decreased as the unit cell

size was scaled down. The largest unit cell size was 7 mm decremented by 0.5 mm until the

smallest unit cell size of 1.5 mm. For the octet-truss in the largest unit cell size, the diameter of

the struts was 2.391 mm, or a volume fraction of 52.6%. For the opti-octet in the largest unit cell

size, the diameters of the struts corresponding to the cubic and BCC struts were 2.843 and 3.242,

respectively. The range of unit cell sizes was intended for the largest to be able to evacuate all

powder down to the smallest where the channels for powder to escape the lattice were expected to

be too small for any powder removal. Each lattice structure sample was a 5x5x5 array of unit

cells. In summary, 72 total samples were created based on the 2 unit cell topologies, 12 unit cell

sizes with the same volume fraction, and 3 replicates of each for repeatability. Figure 5.8 shows a

lineup of 12 as-printed specimens for the octet-truss unit cell topology.

Page 224: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

195

Figure 5.8 As-printed lineup of octet-truss powder removal specimens. The largest specimen is a

5x5x5 array of 7 mm unit cells, approximately 35 mm in width. The smallest specimen is a 5x5x5

array of 1.5 mm unit cells, approximately 7.5 mm in width.

The powder removal specimens where fabricated on an EOS M290 in SS 17-4. Prior to

removal from the build plate, vacuuming, blowing compressed air, bead blasting, and tapping the

plate were all used to attempt to extract all unsintered powder from the lattices. Next, the

specimens were cut off from the build plate using wire electrical discharge machining. After

being removed from the plate, specimens were again individually processed using similar

methods to attempt to remove all unsintered powder.

5.2.4.2 Mass Comparison

For the as-printed mass, the samples were weighed using an A&D GR-202 scale. The

three largest unit cell sizes fell outside of the 210g capacity of the GR-202 scale and were

weighed using an Ohaus Scout II SC4010. Using a density value of 7.77 g/cm3 [371], the as-

designed mass of each specimen was calculated using the as-designed build volume. The relative

error was calculated as the mass difference between the as-printed and as-designed, normalized

Page 225: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

196

by the as-designed value. With the relative error defined in this way, a positive error value

demonstrates excess mass exists within the lattice structure, while a negative error value

demonstrates less mass than the as-designed value.

The comparison of relative error between the octet-truss and the opti-octet is shown in

Figure 5.9. The error bars represent the standard deviation for the three replicates. As a general

trend, as the unit cell size increases, the relative error decreases because as the unit cell size scales

up, the pore size is also scaled up, making it easier for powder to be removed. For the largest unit

cell sizes, there were a few cases where the relative error was negative. The negative relative

error was attributed to the cutoff height from the build plate, where the bottom of each sample

was slightly trimmed, making the mass slightly less than the as-designed model. Also, of note

was that the octet-truss always had a larger relative error than the opti-octet suggesting that more

powder was trapped within the octet-truss.

Page 226: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

197

Figure 5.9 Comparison of the relative error between the as-designed and as-printed mass for the

powder removal specimens. In (B) three regions are predicted based on the relative error (1) no

powder removed, (2) partial removal, and (3) all powder removed, shown with the red ellipses

overlaid.

From the relative error plot in Figure 5.9B, there were three distinct regions that develop

on the plot, (1) a plateau in relative error for the three smallest unit cell sizes, (2) a sloped relative

error region in the middle, and (3) a plateau in relative error for the largest 4-5 unit cell sizes.

This trend is consistent with the explanation that below a certain threshold no powder can be

Page 227: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

198

removed and above a certain threshold all the powder can be removed. For region 1, the relative

error was consistent because all the specimens were essentially solid blocks. For region 3, the

relative error was also similar because there was no powder trapped inside, only slight variations

based on surface roughness or other AM defects. Region 2 represents the transition zone where

the powder can be removed in some cases but is inconsistent. The error bars are the standard

deviation of the averaged data and represent the amount of variation. The largest error bars, or

largest variation, in Region 2 further support the claim in that in the transition region some

powder could be removed but it was inconsistent at the threshold of when powder can be

removed.

Based on the results, the opti-octet demonstrated better powder removal than the octet,

particularly in the transition region. At the extremes of Region 1 and Region 3 the relative error

was expected to be similar between both unit cell topologies because either all powder was

removed or all powder was trapped. However, in the transition region was where the opti-octet

was expected to differentiate itself from the octet-truss. While the octet-truss consistently has a

larger relative error than the opti-octet, the lower and statistically significant relative error of the

opti-octet for unit cell sizes 4.5, 5, 6, 6.5 mm demonstrated improved powder removal compared

to its octet-truss equivalents. One potential confounding effect not taken into account in the mass

measurement comparison was the surface roughness and dross on downskin surfaces. The impact

of surface roughness relative to the amount of total volume could conceivably also follow a

similar trend, where the amount of dross relative to the total volume would be higher for the

smallest lattices and decrease with increasing unit cell size. As such further validation of the opti-

octet was explored through computed tomography (CT) scans to visualize the internal feature of

the lattice structures.

Page 228: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

199

5.2.4.3 Computed Tomography Validation

As further validation of the powder removability study, CT scans were taken of the 1.5,

4.5, and 7 mm unit cell sizes for the octet-truss and opti-octet. Due to the size and density, the

largest sizes were sectioned into 4 vertical columns to obtain a higher scan quality than scanning

the entire specimen as a single piece. The CT scan resolution for the 1.5 mm unit cell size was 15

µm. For the 4.5 mm and 7 mm unit cell sizes the CT scan resolution was 35 µm. Qualitative

comparison of a central slice for each unit cell size and the as-designed model are shown in

Figure 5.10. Though these single slices represent only a fraction of the data, these central slices

were representative of features seen throughout the sample for their respective size and unit cell

topology.

Page 229: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

200

Figure 5.10 Qualitative CT scan comparison of a central slice for the octet-truss and opti-octet unit

cell sizes. The dashed line squares represent a single unit cell within each image. Comparing the 4.5

mm unit cell sizes against the computer generated as-designed cross-section, the octet-truss shows

more trapped powder or partially sintered regions as compared to the opti-octet, validating that the

opti-octet demonstrates improved powder removability.

Page 230: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

201

In comparing each of the three unit cell sizes against the as-designed cross section, the

CT scans confirmed the findings from the mass measurement of the samples shown in Figure 5.9.

In the context of describing the CT scan slices shown in Figure 5.10, the term void refers to an

empty region as detected by the CT scan. The term cavity refers to an internal empty space that

should exist according to the as-designed model. The term channel refers to pathways of

continuous voids in the as-designed model through which unsintered powder would pass to be

evacuated from the sample.

For the 1.5 mm unit cell sizes, the finding from the mass comparison was indicative of no

powder being removed. In nearly every cavity, the CT scans show sintered or trapped powder.

The CT scan also shows small voids located in many of cavity regions, suggesting that some

powder was removed from the sample. However, using Avizo [372] for analysis of the pore

network, or void space mapping, there were no channels large enough to be able to evacuate

unsintered powder from the majority of these internal cavities. This finding suggests that these

voids were not formed by evacuated powder, but occurred during the build, possibly during the

recoat stage or blown out by the air flow across the powder bed. The sintered or trapped powder

in every internal cavity confirms the finding that at the smallest size, all trapped powder either

became sintered or could not be evacuated.

For the 4.5 mm unit cell sizes, the finding from the mass comparison suggested that some

powder was removed from internal cavities but not all, and that the octet-truss had more trapped

powder in its internal cavities and channels than the opti-octet. The CT scans show that for the

octet-truss, many of the cavities were filled with sintered or trapped powder. Cavities near the

outer surface had a greater tendency to be void than cavities near the center of the specimen. This

suggests that it was easier for the powder to find a way out of the specimen from outer cavities

than inner cavities due to the maze of channels to navigate and less risk of lightly sintered powder

blocking the channels. Though only a single slice, this finding was representative of cross

Page 231: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

202

sections throughout the 4.5 mm octet-truss sample. Unfortunately, due to the quality of the scan,

complete segmentation for pore network analysis was not feasible. For the opti-octet, there were a

few cavities that showed deviations from the as-designed cross-section due to sintered powder,

however these were significantly less than the octet-truss. Using pore network analysis in Avizo

for the opti-octet, all void space in the cavities was shown to map to the outside of the unit cell,

demonstrating that powder could be evacuated throughout the unit cell. The sintered or trapped

regions within the octet-truss and lack thereof in the opti-octet, confirm the findings that the opti-

octet shows improved powder removability.

For the 7 mm unit cell size the finding from the mass measurement suggested that these

samples matched very closely with the as-designed model. In some cases, the relative error was

also negative. For the opti-octet cross section, the circles at the bottom of the image are visibly

cutoff, showing the loss in material that accounts for the negative relative error. In addition, due

to the need to section these samples into four vertical columns each, the CT slice shown in Figure

5.10 was slightly offset from the center. Overall, the CT scans show clean cross sections with

channels and cavities clear of unsintered powder, validating the findings that the largest size for

both unit cells were successful in removing all powder from the internal cavities and channels.

Finally, using the as-designed models for the 4.5 mm unit cell size, the tortuosity of each

unit cell topology was computed using pore network analysis in Avizo. Tortuosity is a measure of

how tortuous a path is, in other words, a quantitative measure reflecting how complex the maze is

to remove powder from a structure. In its most basic form, tortuosity is defined as the ratio of

curve length to chord length, meaning a value of 1 is ideal and the more tortuous the path, the

higher the ratio. In the pore network analysis calculates the hydraulic tortuosity. Using a pressure

difference applied across the sample, the absolute permeability of the structure is calculated. With

the flow velocities through each channel known, hydraulic tortuosity is calculated by summing

the length of all velocities and dividing by the sum of the projection of the velocities along the

Page 232: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

203

flow direction. From the pore network analysis, the octet-truss tortuosity ratio was 1.736. On the

other hand, the opti-octet had a tortuosity ratio of 1.419, an 18.3% improvement.

Using mass measurement comparison of the as-designed and as-printed samples,

qualitative CT analysis, and quantitative comparison of the hydraulic tortuosity, the opti-octet

demonstrates improved powder removability over the octet-truss. These results are consistent

with the PRF factors calculated by ALTO which predicted better powder removability for the

opti-octet than the octet-truss. This case study demonstrates an example of using ALTO to create

a novel lattice structure matching the orthogonal effective moduli of a baseline lattice but

exhibiting improved powder removability over the baseline unit cell.

5.3 Case Study 2: Generation of a 3D Unit Cell Topologies with a Fixed Volume Fraction

The purpose of Case Study 2 was to generate several novel unit cells with different

constitutive matrices but similar interconnectivity to a baseline unit cell design. Assuming an

application of a lattice structure with graded stiffness, the intent was to create different stiffness

without changing the volume fraction and while maintaining connectivity between unit cells. For

a given unit cell size and volume fraction, the homogenized constitutive matrix may be

calculated. To adjust the stiffness of a lattice, one typically increases or decreases the volume

fraction, changes the unit cell topology, or a combination of both. Increasing or decreasing the

volume fraction changes the unit cell weight. As a result, a lattice with different unit cells of

varying stiffness generally has a non-uniform weight distribution which may be undesirable for

the application. Changing the unit cell topology within a lattice may lead to connectivity issues

between unit cells. ALTO can be used to generate unit cells with different constitutive matrices

but similar connectivity and volume fraction. In this case study, the goal was to generate three

Page 233: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

204

different unit cell topologies that all have the same volume fraction and similar interconnectivity

but distinct stiffness properties from their homogenized constitutive matrix.

5.3.1 Optimization Description

For this optimization three different objective functions were included: 1) minimization

of error between the target and calculated homogenized constitutive matrix, 2) minimization of

error between the target and calculated volume fraction, and 3) minimization of relevant

elements. This optimization problem also utilized the overlay and DfAM unsupported–N

penalties. A formal description of the optimization problem is given in Eqns. (5.11)-(5.15) with

additional details in Section 3.2.2.2. A summary of the objectives, constraints, and penalties is

shown in Table 5.5 and described in the following paragraphs.

𝑀𝑖𝑛𝑖𝑚𝑖𝑧𝑒 (𝑊𝐻 ,𝑊𝑉𝐹 ,𝑊𝑅𝐸)

𝑆𝑢𝑏𝑗𝑒𝑐𝑡 𝑡𝑜 𝐴𝑒 − 𝐴𝑙𝑜𝑤𝑒𝑟 ≥ 0 (5.11)

𝐴𝑒 − 𝐴𝑢𝑝𝑝𝑒𝑟 ≤ 0 (5.12)

𝐴𝑒 = 𝐴𝑙𝑜𝑤𝑒𝑟 𝑓𝑜𝑟 𝐴𝑒 < 𝐴𝑟𝑒𝑙 (5.13)

𝑊ℎ𝑒𝑟𝑒 𝑊𝑖 = 𝑚𝑖(𝑠𝑖𝑃 + 𝑤𝑖(𝑨, 𝑳)) (5.14)

𝑃 =∑𝑝𝑗(𝑨)

𝑛𝑝

𝑗

, 𝑓𝑜𝑟 𝑗 = 1…𝑛𝑝 (5.15)

Page 234: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

205

Table 5.5 Case Study 2: 3D homogenized unit cells with different stiffnesses optimization summary

Min. Homogenization Error (𝑚𝑖, 𝑠𝑖) = (1,100)

Min. Volume Fraction Error (𝑚𝑖 , 𝑠𝑖, 𝑉𝑓∗) = (1,100, .085)

Min. Relevant Elements (𝑚𝑖, 𝑠𝑖) = (1,100)

Constitutive Scalar (𝑏𝑚) 1

4,1

2, 1

Interconnectivity Nodes See Figure 5.11(E) and (F)

Symmetry XYZ midplanes

𝜃𝑙𝑜𝑤𝑒𝑟, 𝜃𝑢𝑝𝑝𝑒𝑟 (−1, 35) 𝑑𝑒𝑔𝑟𝑒𝑒𝑠

𝑑𝑟𝑒𝑙 0.5𝑚𝑚 𝑑𝑙𝑜𝑤𝑒𝑟, 𝑑𝑢𝑝𝑝𝑒𝑟 (9.5𝑥10−5, 1.5) 𝑚𝑚

Penalties Overlap, Unsupported-N

Borg Evaluations 90,000

Objective 1 was to minimize the total relative error from the target constitutive matrix.

For additional details regarding the homogenization method, see Section 3.2.3.1.3 or

[297,298,361]. The baseline unit cell (Dbase) used to determine the target constitutive matrices

was a diamond unit cell, unit cell size 8 mm and strut diameters 1 mm; see Figure 5.11A and

Figure 5.11B. The diamond lattice structure was chosen for its common use in other lattice

structure research. The graph representation showing the locations of the unit cell struts is shown

in Figure 5.11C and Figure 5.11D. After calculating the Dbase homogenized constitutive matrix,

�̃�𝐷𝐻, shown in Eqn. (5.16), the target constitutive matrices were calculated as scalar multiples

using Eqn. (5.17), where 𝐸 is the material Young’s modulus and the 𝑚-th target constitutive

matrix is scaled by 𝑏𝑚. For this case study, three optimizations were set up 𝑚 = 1,2,3 for 𝑏𝑚 =

1

4,1

2, 1, respectively. Finally, the objective function for objective 1 is shown in Eqn. (5.18).

�̃�𝑫𝑯 = 𝐸

[ 𝟗. 𝟓 𝟗. 𝟓 𝟗. 𝟓 𝟎 𝟎 𝟎

𝟗. 𝟓 𝟗. 𝟓 𝟎 𝟎 𝟎𝟗. 𝟓 𝟎 𝟎 𝟎

𝟎 𝟎 𝟎𝟎 𝟎

𝒔𝒚𝒎 𝟎]

𝑥10−3 (5.16)

�̃�𝑚∗ = 𝑏𝑚�̃�𝐷

𝐻 (5.17)

Page 235: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

206

𝑤𝐻 = ∑

(𝐸𝑘𝑙∗ − 𝐸𝑘𝑙

𝐻)

𝐸𝑘𝑙∗ 𝑓𝑜𝑟 𝑘, 𝑙 = 1…6 (5.18)

Objective 2 was to minimize the relative error between the calculated and target volume

fraction; see Eqn. (5.19). The target volume fraction was 8.5%, the same as the volume fraction

of Dbase. Objective 3 was to minimize the number of relevant elements to reduce complexity of

the unit cell; see Eqn. (5.20).

𝑤𝑉𝐹 =|𝑉𝑓

∗ −𝑨𝑇𝑳𝑉𝑢𝑐

|

𝑉𝑓∗

(5.19)

𝑤𝐴𝐸 = 𝑐𝑜𝑢𝑛𝑡(𝐴𝑒 > 𝐴𝑟𝑒𝑙) (5.20)

Beginning with a 5x5x5 cubic array of nodes, each node was connected to every other

node in the structure, resulting in 7750 elements or design variables. Next, the optimization

constraints were the relevant diameter limit and bounds on the upper and lower limit of the cross-

sectional area. Ground structure restrictions were limits on the interconnectivity nodes based on

Dbase (see Figure 5.11), surface elements, build angle, and symmetry. Cubic symmetry across

the XYZ mid-planes was applied such that the interconnectivity restriction was maintained. The

optimization penalties were overlap and unsupported‒N. After application of the ground structure

restrictions, the total number of design variables was reduced from 7750 to 67. Figure 5.11E and

Figure 5.11F show the side and isometric views of the initial ground structure for the

optimization, where each line represents a truss element.

Page 236: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

207

Figure 5.11 Case Study 2 is a multi-objective optimization for creating a lattice structure that mimics

the homogenized constitutive matrix of a Diamond unit cell (A) and (B). The unit cell graph is shown

in (C) and (D). To ensure similar interconnectivity to the Diamond unit cell, the full ground structure

is reduced to the initial configuration shown in (E) and (F).

5.3.2 Optimization Results

The results of the case study were three unique cells with homogenization matrices

matching those of the target constitutive matrices. Examples of these unit cells are shown in

Figure 5.12. The unit cells in each solution were found to be based on a 4-element pattern, with

the best shown in Figure 5.12C. This four-element configuration is referred to as a tetrahedra.

Page 237: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

208

Figure 5.12 Case Study 2 resulted in three unit cells which match the scaled target constitutive

matrix of the diamond unit cell. In each solution, a fundamental tetrahedra (grey) shape was found

with decreasing diameter. For the middle and top solutions, several non-load bearing elements (blue)

form a mechanism around the tetrahedra shape, based on the pin-joint assumption, to

simultaneously match the volume fraction and homogenization targets.

As shown in Figure 5.12, each optimization found a variation of a simple tetrahedra

pattern that exactly mimics the diamond unit cell constitutive matrix properties. This tetrahedra

fulfills two of the three objectives by matching the target constitutive matrix and minimizing the

number of elements to only four. To achieve the final objective, matching the volume fraction of

the diamond unit cell, the optimization added elements that contribute to the weight but not the

Page 238: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

209

stiffness, shown in a blue color in Figure 5.12. That is, these elements do not change the

homogenized constitutive matrix because they rely on the pin joint assumptions at the nodes,

meaning that the elements act like a mechanism and do not increase the stiffness of the unit cell.

From here, each design could be further refined through designer interpretation to more

accurately represent a hinged mechanism. It is interesting to note that the solution shown in

Figure 5.12B has a ring of elements that surround the tetrahedra, never connecting. This ring was

formed as additional weight to match the volume fraction and by design interpretation could be

made to bear very little load if the elements were reduced in size near the upper and lower node.

In the end, both the 𝑏𝑚 =1

4,1

2 solutions represent a load bearing tetrahedra with mechanism-like

elements for adding weight without stiffness.

Regarding the 1x stiffness cell result, Figure 5.12C, the solution is referred to as the

tetrahedra unit cell, which when patterned forms an interlocked lattice. As compared to the

diamond unit cell, the tetrahedra represents a more scalable solution because the smallest feature

size is larger than that of the diamond unit cell. In addition, powder removal from the interlocked

lattice is expected to be easier due to the larger spacing between lattice struts than the diamond

unit cell. On the other hand, the longer struts are likely to make the tetrahedra unit cell more

prone to buckling.

By comparison of both the tetrahedra and diamond unit cells, the tetrahedra shape is

actually a small section of the original diamond structure. Upon further inspection of the diamond

unit cell, one can see that the tetrahedra is repeated 4 times within the unit cell; see Figure 5.13.

From this finding, it may be argued that both structures are the same. However, the diamond and

interlocked lattice structures are indeed different because the repetition or patterning of this unit

cell results in a different structure; see Figure 5.13. In the diamond lattice, the tetrahedra is

repeated in a checkerboard pattern as shown in the patterned lattice schematic. For the interlocked

lattice, the repetition is continuous; see Figure 5.13. The continuous patterning allows for the

Page 239: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

210

same predicted constitutive matrix while using larger element diameters and fewer elements than

the diamond lattice. Due to objective 4, the minimization of relevant elements, fewer and thicker

elements mean that the structure is better for powder removal than the diamond lattice and can be

more scalable for printing than the diamond lattice because the smallest diameters of the

tetrahedra were larger than the diameters for the diamond unit cell.

Figure 5.13 The interlocked lattice is a unique patterning of the tetrahedra base that results in a

scaled and interlocked set of two diamond lattice. In the final patterned interlocked lattice, the unit

cells only connect to diagonally neighboring unit cells.

One unique result was that the interlocked lattice was named such because after

patterning, it formed two interlocking lattices that move independent of each other, much like

interconnected chain mail. Figure 5.14 shows the interlocked lattice, with the blue lattice being

independent from the grey. The unit cells connect to diagonally neighboring unit cells, as opposed

to connecting to other unit cells neighboring the 6 cubic faces.

Page 240: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

211

Figure 5.14 The Interlocked lattice is a set of two interconnected lattices. The upper lattices were

printed in Alumide® on an EOS P110. The lower left image shows the interlocking lattices,

distinguished by the blue and grey colors.

Case Study 2 demonstrates the ability to create novel unit cell topologies with specified

properties and interconnectivity to neighboring unit cells. For a graded lattice stiffness, changing

unit cell topology or reducing the volume fraction within a lattice can lead to issues of

interconnectivity between unit cells or varying weight distribution. By fixing the volume fraction

and interconnectivity between lattice structures, ALTO manipulates the unit cell topology to vary

the lattice stiffness while enabling a uniform weight distribution. The three solutions from Case

Study 2 represent a novel approach to tailoring lattice stiffness while simplifying

interconnectivity for the same weight. The optimization accomplishes this by combining

mechanism-like behavior into novel lattice structure designs. Using the DfAM objective function

minimization of relevant elements, a novel tetrahedra unit cell was discovered which offers the

same theoretical constitutive matrix as the diamond lattice and larger spacing than the diamond

Page 241: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

212

lattice for powder removability. In addition, the tetrahedra unit cell has greater scalability because

the diameter of the struts is larger than the diameter of the diamond unit cell struts.

5.3.3 Fabrication and Compression Testing

For demonstration of printing the diamond and interlocked lattice structures, a set of

diamond and interlocked lattices were and evaluated using uniaxial compression testing. The test

specimens were built on the same build plate and follow the same specimen format and testing

procedure as in Section 5.2.3.1. The strut diameters for the diamond were a uniform thickness of

0.5 mm. For the interlocked lattice, the 1x stiffness optimized result was printed using a uniform

diameter 0.707 mm.

The average stress-strain curves from four samples each for the diamond and interlocked

lattices are shown in Figure 5.15. The grey band for each curve represents the average plus and

minus one standard deviation of the experimental data. The peak load for each lattice was

determined to be 9.2 ± 0.03 MPa and 5.9 ± 0.10 MPa for the diamond and interlocked lattices,

respectively. Based on the stress strain curves, the linear region was determined to be between

7.5-60% of the peak load. In Figure 5.15, the effective modulus region for each lattice is denoted

by the solid portion of the stress strain curves. From this region, the effective modulus of the

lattice was calculated using a linear least squares fit. The diamond and interlocked lattice

effective moduli were calculated to be 305 ± 4.3 MPa and 162 ± 6.8 MPa, respectively.

Page 242: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

213

Figure 5.15 Average stress strain curves for the diamond and interlocked lattices. The solid portion

of the curves represents the region where the effective modulus was measured, 7.5-60% of the peak

stress. The grey band behind each curve represents the average plus and minus the standard

deviation of the experimental data.

Based on the need to relax the boundary condition discussed for the previous case study,

the ground structure formulation in ALTO is not capable of accurately capturing the effective

moduli of the diamond and interlocked lattices. With the “relaxed” boundary conditions as

described in Section 5.2.3.3, the diamond and interlocked lattices become mechanisms. For

mechanism-like behavior, the effective modulus captured during experimental compression tests

is related directly to the stiffness at the nodes as opposed to axial stiffness in the struts. With a

significantly higher number of nodes in the diamond structure than the interlocked, it is not

surprising that the effective modulus is higher in the diamond lattice than the interlocked lattice

when comparing the results for “relaxed” boundary conditions measured experimentally. For a

ground structure model composed of truss elements, the pin joints provide no stiffness and a zero-

stiffness result is calculated in the homogenized constitutive matrix from ALTO. However, with a

raw material Young’s modulus of 115 GPa, the experimental results further support the need to

Page 243: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

214

relax the boundary conditions, with the extremely low calculated effective modulus values, only

0.27% and 0.14% when normalized by the raw material Young’s modulus.

5.4 Case Study 3: Generation of a 2D Orthotropic Unit Cell Topology

Up to this point, the case studies have been for isotropic materials, materials which

exhibit the same effective mechanical properties in the three primary orthogonal directions.

ALTO theoretically has the ability to match any constitutive matrix. Case Study 3 is a 2D

example of generating a unit cell to match an orthotropic target constitutive matrix. The target

constitutive matrix was selected so that the effective modulus of unit cell was three times larger in

the X-direction than the Y-direction; see Figure 5.16.

Figure 5.16 The purpose of Case Study 3 is to generate a unit cell topology that has an orthotropic

constitutive matrix, specifically that it is three times stiffer in the X-direction than in the Y-direction.

5.4.1 Optimization Description

For the optimization, a single objective function was utilized, minimization of error

between the target and calculated homogenized constitutive matrix. This case study used the

gradient-based DfAM penalty for minimizing complexity and the DfAM relevant diameter limit

constraint. The optimization was performed using the gradient-based algorithm fmincon with a

volume fraction constraint on the total volume. A formal description of the optimization is shown

Page 244: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

215

in Eqns. (5.21)-(5.27). A summary of the optimization problem is shown in Table 5.6. For more

details on the optimization formulation or objective, constraint, and penalty functions see Chapter

3.

𝑀𝑖𝑛𝑖𝑚𝑖𝑧𝑒 𝑓(𝑊, 𝑃)

𝑆𝑢𝑏𝑗𝑒𝑐𝑡 𝑡𝑜 𝐴𝑒 − 𝐴𝑙𝑜𝑤𝑒𝑟 ≥ 0 (5.21)

𝐴𝑒 − 𝐴𝑢𝑝𝑝𝑒𝑟 ≤ 0 (5.22)

𝑨𝑇𝑳 − 𝑉𝑓∗𝑉𝑚𝑎𝑥 ≤ 0 (5.23)

𝐴𝑒 = 𝐴𝑙𝑜𝑤𝑒𝑟 𝑓𝑜𝑟 𝐴𝑒 < 𝐴𝑟𝑒𝑙 (5.24)

𝑊ℎ𝑒𝑟𝑒 𝑓 = (𝑊)(𝑃) (5.25)

𝑊 = 𝑐𝐻𝑚𝐻𝑤𝐻(𝑨, 𝑳) (5.26)

𝑃 = 𝑝(𝑨, 𝑳) (5.27)

Table 5.6 Case Study 3: 2D orthotropic unit cell generation optimization summary

Min. Homogenization Error 𝑚𝑖 = 1

Volume Fraction Constraint 𝑉𝑓∗ = 0.15

Material Modulus Ratio (𝛼) 3

Poisson’s Ratio (𝜈12) 0.6

Shear Modulus (𝐺12) 𝐸𝑥𝑥

3(1 + 𝜈𝑥𝑦)

Target Constitutive Matrix (�̃�∗) 𝐸𝑥𝑥 [1.136 0.227 0

0.379 0𝑠𝑦𝑚 0.208

]

DfAM penalty function parameters (𝛼1, 𝛼2, 𝛼3, 𝛼4) = (2, 100,100

99,98

99)

DfAM penalty function levels 𝛾 = 0,1

6,1

3,1

2,2

3

𝑑𝑟𝑒𝑙 0.3, 0.5, 1.0 mm

𝑑𝑙𝑜𝑤𝑒𝑟, 𝑑𝑢𝑝𝑝𝑒𝑟 (1𝑥10−6, 2) mm

𝜃𝑙𝑜𝑤𝑒𝑟, 𝜃𝑢𝑝𝑝𝑒𝑟 (1, 35) degrees

Symmetry XY midplanes

The design space for the optimization problem was a 5 mm square unit cell discretized by

a 4x4 array of nodes. The objective for the optimization was to minimize the error between the

target constitutive matrix and the calculated homogenized constitutive matrix. In this formulation,

the solution was restricted to 2D orthotropic materials, resulting in a [3x3] symmetric constitutive

matrix defined by two elastic moduli, two Poisson’s ratios, and the shear modulus shown in Eqn.

Page 245: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

216

(5.28), where �̃� is the constitutive matrix, 𝐸𝑥𝑥 and 𝐸𝑦𝑦 are elastic moduli, 𝜈𝑥𝑦 and 𝜈𝑦𝑥 are

Poission’s ratios, and 𝐺𝑥𝑦 is the shear modulus for a plane stress condition. Using symmetry, the

terms in the 𝐸12 and 𝐸21 positions are equal, resulting in Eqn. (5.29), where 𝛼 is the ratio of the

moduli (Exx

Eyy) and Poisson’s ratios (

𝜈𝑥𝑦

𝜈𝑦𝑥), resulting in a constitutive matrix defined by four

parameters (𝐸𝑥𝑥 , 𝜈𝑥𝑦 , 𝐺𝑥𝑦, 𝛼). According to the problem statement 𝑎 = 3 so that the effective

modulus in the X-direction is three times larger than in the Y-direction. In this case, the focus of

the constitutive matrix was the effective modulus in the X- and Y-directions so the value of 𝜈𝑥𝑦

was chosen arbitrarily to be 0.6. In summary, by setting (𝜈𝑥𝑦, 𝐺𝑥𝑦, 𝛼) = (. 6,𝐸𝑥𝑥

3(1+𝜈𝑥𝑦), 3) the final

target constitutive matrix was determined, normalized by 𝐸𝑥𝑥; see Eqn. (5.30). The objective

function for the optimization is shown in Eqn. (5.31).

�̃� =

[

𝐸𝑥𝑥1 − 𝜈𝑥𝑦𝜈𝑦𝑥

𝜈𝑦𝑥𝐸𝑥𝑥

1 − 𝜈𝑥𝑦𝜈𝑦𝑥0

𝜈𝑥𝑦𝐸𝑦𝑦

1 − 𝜈𝑥𝑦𝜈𝑦𝑥

𝐸𝑦𝑦

1 − 𝜈𝑥𝑦𝜈𝑦𝑥0

0 0 𝐺𝑥𝑦]

(5.28)

�̃� =1

𝛼 − 𝜈𝑥𝑦2 [

𝛼𝐸𝑥𝑥 𝜈𝑥𝑦𝐸𝑥𝑥 0

𝐸𝑥𝑥 0

𝑠𝑦𝑚 (𝛼 − 𝜈𝑥𝑦2 )𝐺𝑥𝑦

] (5.29)

�̃�∗ = 𝐸𝑥𝑥 [

1.136 0.227 00.379 0

𝑠𝑦𝑚 0.208] (5.30)

𝑤𝐻 =

∑|𝐸𝑘𝑙∗ − 𝐸𝑘𝑙

𝐻 |

𝐸𝑘𝑙∗ 𝑓𝑜𝑟 𝑘, 𝑙 = 1,2,3 (5.31)

For optimization constraints, the case study has a limit on the upper and lower bounds of

the element cross-sectional areas, a relevant diameter limit, and a total volume fraction limit.

Ground structure restrictions include the build angle limit and symmetry across the horizontal and

Page 246: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

217

vertical midplanes. The restrictions reduced the number of design variables from 120 to 30.

Depending on the overall size of the unit cell, high complexity within the unit cell could result in

trapped powder or fused members. A DfAM penalty function to reduce unit complexity was

applied for the gradient-based algorithm. The function penalizes elements near the relevant

diameter limit, encouraging fewer and thicker elements; see additional details in Section

3.2.3.3.2. The penalty function was controlled by four parameters and then scaled by an exponent,

the penalty function level, to change the level of impact from the penalty function. The

parameters chosen for this optimization are given in Table 5.6 and an image for the resulting

penalty function is shown in Figure 5.17. In addition to the penalty function to reduce complexity,

increasing the relevant diameter limit forces fewer members due to the volume constraint. This

effectively reduces the total number of elements that can be used to form the unit cell. Because

the gradient-based algorithms are prone to find local minima, multiple initial guesses were used

to search the design space. DfAM considerations applied to this case study were the build angle

limit, relevant diameter limit, and penalty function to reduce overall unit cell complexity.

Figure 5.17 Graph of the DfAM penalty function that penalizes elements near the relevant cross-

sectional area limit and reduces non-critical elements. Image adapted from [361].

Page 247: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

218

5.4.2 Optimization Results

To explore the penalty function and relevant diameter limit, a range of values were used

for each. For each penalty function level (𝛾 = 0,1

6,1

3,1

2,2

3) and relevant diameter limit

(𝑑𝑟𝑒𝑙 = 0.3,0.5,1.0), the optimization was initialized from 20 different randomly generated initial

guesses. The range of penalty function levels allows for varied impact from the penalty function.

For 𝛾 = 0, there was no penalty factor. With each increasing penalty factor level, the impact of

the penalty increases, providing a range of solutions. From 20 different initial guesses for each

combination of penalty function level and relevant diameter limit, the results represent a total of

300 different instances of the optimization problem. The total computation wall time was

approximately 3 hours using 6 cores. Figure 5.18 shows the best solutions from 20 randomized

initial guesses, each for a different penalty factor level and relevant diameter limit. Though there

were similar features between many of these designs, the figure shows that there were a variety of

solutions that closely matched the target constitutive matrix. The penalty function and relevant

diameter limit may be used to further refine the solution set and promote or penalize designs with

characteristics that are beneficial or challenging for AM processes.

Page 248: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

219

Figure 5.18 Lowest optimization scores for the constitutive matching objective function. Each image

represents the best solution of 20 initial guesses for the particular combination of penalty factor level

and relevant diameter limit. Image adapted from [361].

In this case study, both the penalty function and relevant diameter limit were used to

encourage simpler solutions by encouraging elements cross-sectional area larger than the relevant

diameter limit, thereby reducing the total number of relevant elements. In general, viewing Figure

5.18, as you look from left to right (increasing relevant diameter limit) or from top to bottom

(increasing penalty factor level), the complexity of the solution decreases, showing fewer and

thicker members within the unit cell. While both the penalty and relevant diameter were

successful in reducing the complexity of the unit cell topology, this comes at a cost of error in the

homogenized constitutive matrix because there were fewer elements in a unit cell and tighter

Page 249: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

220

limits on the diameters of each element. For example, of the 20 random initial guesses for 𝛾 =

0 , 𝑑𝑟𝑒𝑙 = 0.3, 95% of the designs had a total error in the constitutive matrix of less than 10%.

However, for 𝛾 =2

3, 𝑑𝑟𝑒𝑙 = 0.3 only 25% of designs had an error of less than 10%. For 𝑑𝑟𝑒𝑙=1.0,

only 3 solutions out of out all 5 penalty function levels with 20 initial guesses each (100

optimizations in total) were able to achieve a total error of less than 10%. Therefore, reducing the

complexity of the unit cell was desirable but it may come at the expense of increased error for

matching the target constitutive matrix.

Another method to reduce cell complexity was to reduce the number of elements based

on the node configuration or the interconnectivity ground structure restriction. For example,

setting the cross-sectional area of elements connected to the corners to 𝐴𝑙𝑜𝑤𝑒𝑟 forces those

elements to be uneconomical and reduces the number of design variables to only 18. Figure 5.19

shows an example of a unit cell generated with 𝑑𝑟𝑒𝑙 = 0.5 and 𝛾 =1

3 and no elements connected

to the corners. The deformed shape of the unit cell is shown under the same compression load

applied horizontally Figure 5.19A and vertically Figure 5.19B. As shown in the figure, the

deformation was three times larger in the vertical direction than it was in the horizontal direction,

demonstrating the target constitutive matrix. Figure 5.20 shows the patterned lattice structure

based on the unit cell topology from Figure 5.19.

Page 250: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

221

Figure 5.19 Optimized unit cell topology with the area of the elements connected to the corners set to

the lower limit. Compression in X and Y is also shown, demonstrating the target constitutive matrix.

Total error in the constitutive matrix was 3x10-6 %. Image adapted from [361].

Figure 5.20: Example of a lattice structure patterned from the unit cell topology shown in Figure

5.19. This unit cell topology was generated using a penalty function level of one third and the relevant

diameter limit set to 0.5 mm. Image adapted from [361].

A) X-Compression

B) Y-Compression

Y

X

�̃�𝑯 = 𝐸𝑥𝑥 [1.136 0.227 0⬚ 0.379 0𝑠𝑦𝑚 ⬚ 0.208

]

Page 251: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

222

Case Study 3 demonstrates the ability of ALTO to successfully generate unit cell

topologies that match an orthotropic constitutive matrix. The DfAM considerations and penalty

function reduced complexity and improved manufacturability of the lattice structure. With this

approach, novel metamaterials may be developed and customized for specific applications

through optimization.

5.5 Discussion of Results

While the results for each case study show generation of novel unit cell topologies,

topology optimization through an approach like ALTO is not without its limitations or challenges.

The focus of this discussion section is on results found from each of the case studies. Excluded

from this discussion are many of the general limitations and challenges associated with ground

structures, which were covered previously in Section 4.4.

Case Study 1 demonstrates a novel powder removability objective function for improved

manufacturability. Using the PRF, the opti-octet unit cell topology was generated and

demonstrated improved powder removal over the baseline octet-truss unit cell. Aside from

generation of the novel unit cell, this case study also included experimental validation of ALTO

through compression testing of the two lattice structures. Two AM modifications were considered

and discussed in this section including a novel build angle dependent effective cross-sectional

area to account for the staircase effect. The ALTO homogenization calculation was found to be an

accurate prediction of the compressive effective modulus when appropriate boundary conditions

were considered. One limitation of the homogenization approach was the boundary conditions, as

discussed in Section 5.2.3.2. For lattice structures, it has been shown in other research that the

number of layers can significantly change the experimentally calculated effective properties

[207]. In the work by Lei et al. [207], experimental data and finite element analysis are from

Page 252: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

223

compression testing of lattice samples with 1, 3, 5, and 7 layers, the theoretical values are

calculated based on single unit cell for which the number of layers is cannot be taken into

account. Table 5.7 shows data from Lei et al. [207] for two different lattice types that further

supports the changing boundary conditions for lattices with different numbers of layers.

Table 5.7 Lattice structures demonstrate a substantial difference in mechanical properties based on

the number of layers. As the number of layers increase, the effective modulus decreases, data from

[207].

Lattice Data Type 1 Layer 3 Layers 5 Layers 7 Layers

BCC Theoretical 225.00 225.00 225.00 225.00

Experimental 198.62 ± 2.36 87.11 ± 1.24 39.16 ± 1.07 21.71 ± 0.86

FEA 210.79 109.68 41.79 18.11

BCCZ Theoretical 554.90 554.90 554.90 554.90

Experimental 490.22 ± 3.41 378.32 ± 2.87 330.31 ± 2.66 317.07 ± 2.05

FEA 557.87 443.26 391.40 366.70

A challenge with lattice structures is that depending on how they are applied within a

structure they may exhibit very different performance. For example, consider a component that

has an embedded lattice for lightweighting and is surrounded by solid material. In this case, the

lattice would be highly constrained by the solid material surrounding it and exhibit a high

stiffness. However, consider a second component in which the lattice is not surrounded by solid

material. In this second case, if the lattice is left unconstrained by a surrounding structure, the

stiffness of the lattice would be significantly less than in the first component. While other

traditional materials may exhibit a similar behavior, the result may be exaggerated for lattices

because the stiffness of bending at the nodes is significantly less than the axial stiffness along the

truss element. In unconstrained cases, the bending or deformation near the nodes significantly

reduces its stiffness. Despite this challenge, the effective mechanical properties can be accurately

predicted if the boundary conditions are carefully considered. As with the example of Case Study

1, the effective modulus predicted in ALTO was found to match within 6% to the experiment

data.

Page 253: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

224

One other limitation demonstrated with ALTO was the ability to capture geometric

variations based on the AM manufacturing. In this case, AM manufacturing modifications were

considered for correcting the ground structure prediction to match the as-printed geometry of the

lattice structure. These corrections were for the staircasing effect and reduction in the cross

section based on the laser scan strategy. The two corrections showed promise in improving the

predicted result but other factors such as the 1) dross on horizontal elements or 2) geometric

anisotropy due to the build process were not accounted for. For example, on horizontal elements

it is known that a circular cross-section becomes ellipsoidal due to the extra material that is

picked up on the downskin surface. At the bottommost extreme of this ellipse, the powder is

likely lightly sintered and would not significantly contribute to the stiffness, meaning the entire

ellipse cross section is not contributing to the stiffness. However, some of the added material is

likely to make a contribution to the stiffness. The question then arises, should this correction and

prediction be performed as part of the topology optimization step or corrected as the part is sliced

model in preparation for printing to more accurately represented the as-designed geometry.

Likely there is a balance of both steps that needs to occur.

A second issue mentioned is the anisotropic behavior of printed lattice structures,

specifically due to geometric discrepancies in the as-designed versus as-printed topology, as

opposed to material anisotropy due to AM. The dross on horizontal elements, mentioned earlier,

means that for elements that were printed vertically, the cross-section should circular compared to

the elliptical cross-section of horizontal elements. As a result, the AM process geometrically

induces anisotropy when compared to the original as-designed model. For example, in this work,

the primary load path for compression testing of the opti-octet would be along the four vertical

columns with the circular cross-sections. However, if tested in X- or Y-axis compression, the

result is likely to vary due to the primary load path now along the ellipsoidal cross-sections.

While these geometric variations due to surface roughness or dross occur in large non-lattice parts

Page 254: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

225

as well, the issue is likely to be exaggerated for lattices because the surface roughness is much

closer to the lattice feature sizes than in a non-lattice part.

In Case Study 2, the optimization was successful in finding solutions that met the desired

criteria. However, early work on this case study highlighted several potential issues that required

additions or corrections in ALTO for 3D unit cell topologies. Two examples of these were

overlapping and disconnected material. Overlapping or intersecting elements at non-nodal

locations was common in many cases because there were so many elements crossing in the unit

cell. Due to the discrete nature of the optimization, it determines the effective mechanical

properties by beams connected to the array of nodes. If elements intersect at non-nodal locations,

ALTO sees these as separate elements, essentially sliding past each other, as opposed the

inevitable fusion that would occur during printing. As a result, this gave rise to the penalty

function which discourages designs with overlapping elements in the solution, resulting in more

print-ready solutions and less design interpretation.

Disconnected material was an issue that arose in the 𝑏𝑚 =1

4,1

2 examples. In these cases,

the optimization was trying to increase the volume fraction without changing the constitutive

matrix. One of the simplest ways to accomplish this was to include elements that never connect to

the load bearing structure, thereby increasing volume fraction without stiffening the structure. As

opposed to simply requiring each individual element to meet the build angle requirements, the

support structure penalties incorporate a more rigorous approach to ensure each element is

buildable.

Without the DfAM build angle restriction or unsupported‒N penalty applied to Case

Study 2, the solutions would not be feasible to print. Figure 5.21 shows a series of results that

occur without DfAM restrictions and penalties. Due to the addition of DfAM restrictions, such as

the build angle, overlap penalty, and unsupported penalties, the optimization resulted in more

print-ready solutions requiring less design interpretation.

Page 255: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

226

Figure 5.21 Without DfAM restrictions and penalties to ensure manufacturability, there were a

variety of pitfalls that appear in optimization solutions. The DfAM considerations help to reduce the

need for design interpretation and prepare more print-ready solutions.

Finally, for Case Study 3, the result demonstrated a method to reduce unit cell complexity

for DfAM using a gradient-based penalty function; however, a major challenge continues to be

the ability to reduce complexity and match a target constitutive matrix. With such a variety of

solutions that were able to closely match the constitutive matrix, there were numerous local

minima where the optimization became trapped. The penalty function was effective in reducing

Page 256: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

227

complexity of the unit cell, however, only to an extent due to the tradeoff with the constitutive

matrix error.

One of the primary challenges or limitations with Case Study 3 was the repeatability of

gradient-based algorithms. In this type of problem there were many different solutions that could

match the target constitutive matrix. For a given starting ground structure, the algorithm was

consistent in finding the same results; however, it was clear that numerous local minima existed

in the solution space because the 20 different randomized starting guesses frequently resulted in a

variety of different solutions. This result was caused by the single search direction nature of

gradient-based algorithms versus the multiple search paths of evolutionary algorithms. In general,

gradient-based optimization case studies were found to require significantly more time upfront

tweaking parameters and adjusting the formulation. The parameters and tweaking were problem

specific and have to be repeated with each type of optimization. If ALTO were consistently

solving the same types of problems, the formulation could likely be tuned in for a specific set of

objectives and constraints and take advantage of the faster speed of a single search direction for

gradient based-optimization algorithms as compared to evolutionary algorithms. In contrast, case

studies utilizing evolutionary algorithms were found to be more robust to varying applications

and less affected by tweaking different parameters, at the expense of potentially increased

optimization time. As the intent of ALTO is a systematic optimization method for a variety of

unit cell topologies and an application agnostic purpose, evolutionary algorithms are

recommended. The ability of evolutionary algorithms to handle multiple objectives and explore

multiple search directions was found to be amendable to unit cell topology optimization.

Page 257: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

228

5.6 Closing Remarks

The focus of this chapter is the use of ALTO for generating unit cell topologies that

match a target constitutive matrix. Three different case studies were presented showing multiple

examples of uses for ALTO in generating novel and customized unit cell topologies. Case Study

1 demonstrates matching the orthogonal effective moduli of a baseline unit cell topology while

showing improved powder removability. In addition, this case study shows compression testing

data to validate the ground structure model. Case Study 2 shows a unique exploration of

generating unit cell topologies with decreasing stiffness while maintaining the same target

volume fraction, forcing the volume to be rearranged, not removed, from the unit cell. Finally,

Case Study 3 shows generation of a unique 2D unit cell that exhibits an orthotropic material

constitutive matrix. In all cases, DfAM considerations increase manufacturability of the solution

to ensure that the optimized solutions can be fabricated with minimal design interpretation and no

support material. AM is a powerful manufacturing process that enables advanced design and

complexity to be incorporated through lattice structures and it requires development of systematic

topology optimization approaches like ALTO to generate novel unit cell topologies and ensure

manufacturability.

Page 258: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

229

Chapter 6 LUCIE: Lattice Unit-cell Characterization Interface for Engineers

The focus of this chapter will be on the development of a comprehensive compilation of

mechanical properties for lattice structures with regular unit cell topologies through an interactive

interface. The chapter includes a brief summary of the relevant background and motivation, the

methods for gathering sources and extracting data, developing a graphical user interface LUCIE

(Lattice Unit-cell Characterization Interface for Engineers), and finally three examples of how

LUCIE can be used to understand lattice structures. This project was completed with help from

undergraduate student Joseph Berthel. From this work, LUCIE contains 69 sources which

resulted in over 1400 experimental data points, over 200 finite element data points, and over 45

analytical models for visualization of lattice structure. It is the largest known database for

mechanical property data for AM lattice structures [373]. The following objectives and sub tasks

are addressed in this chapter:

5. Develop mechanical property database for metal lattice structures from current literature

containing analytical, finite element, and experimental data

5.1. Perform literature search for metal lattice structure data

5.2. Collect data from selected papers

5.3. Create user interface for generating mechanical property charts of collected data

6.1 Background and Motivation

Lattice structures are frequently designed into a component through standard computer-

aided design (CAD) modeling software (e.g., SolidWorks [374], Autodesk Inventor [375], PTC

Creo [376]) or AM-specific software (e.g., Autodesk Netfabb Ultimate [126], ANSYS

SpaceClaim [238], nTopology nTop Platform [235], 3D Systems 3DXpert [237], Materialise 3-

matic [377]). Within the AM-specific software, a library of unit cell topologies to generate

Page 259: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

230

different lattice structures or infill patterns are offered; however, their implications on the bulk

(i.e. macroscale) mechanical properties of the component are unknown in many cases (e.g., the

unit cell topology has inherent mechanical properties in compression and shear that will affect the

overall component performance).

Various methodologies have been recommended regarding the design or inclusion of

lattice structures in a component [228,334]. Tamburrino et al. [334] discuss unit cell selection

based on the choice of either a stretched-dominated or a bending-dominated lattice structure

[334]. However, grouping all truss-based lattice structures into these two categories without

distinguishing them further essentially leaves the decision wide open. In recent work by Bhate

[228], the first of four questions regarding cellular material design is, “What is (are) the optimum

unit cell(s)?” [228]. Bhate describes several methods that have been explored in the literature,

namely, Maxwell’s stability criterion, Gibson-Ashby models of cellular materials, computational

analysis, and experimental testing.

In a recent publication, Maconachie et al. [250] provide a review of topics surrounding L-

PBF lattice structures. The topics include types of lattice structures and their applications,

processing parameters, microstructure and mechanical properties, numerical modeling,

experimental approaches, and outstanding challenges. As discussed in the review, with all the

current research into lattice structures, there is a need to compile and provide a comprehensive

analysis of the data available. This review does compile some experimental data; nonetheless, a

comprehensive compilation of metal AM lattice structure performance data in literature does not

exist.

To expand the current understanding of lattice structure properties and to improve unit

cell selection, this work seeks to compile reported mechanical properties for AM lattice structures

of regular unit cell topologies, gathered through analytical modeling, finite element analysis

(FEA), and experimental testing. The gathered data was compiled and illustrated in Ashby-style

Page 260: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

231

plots through the Lattice Unit-cell Characterization Interface for Engineering (LUCIE), a

graphical user interface that allows researchers to easily compare mechanical properties of metal

lattice structures based on different unit cell topologies made with AM. Similar to traditional

material selection through Ashby plots [378], this work seeks to make lattice structure selection

available through scatter plots with axes chosen based on lattice structure properties. The primary

intent was not to provide exact mechanical property values, but instead to provide the designer

with visualization of performance trends based on the unit cell topology.

6.2 Methods

In this section, the methods for identifying relevant research articles, gathering and

storing data, visualizing the data, as well as assumptions made during data collection are

described. A graphic summary of the methods section is shown in Figure 6.1

Page 261: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

232

Figure 6.1 Graphic summary of the methods section including how sources were found and refined,

data collection and assumptions made, and finally incorporation into LUCIE.

6.2.1 Literature Compilation

To identify research articles containing relevant lattice structure data, literature searches

were conducted in multiple databases such as Compendex. The primary goal of the search was to

find research articles from theses, dissertations, conference, and refereed journal articles

containing mechanical performance information for metal lattice structures made with AM

powder bed fusion (laser or electron beam).

With the recent increase in research in AM, there are a wide variety of terms to describe

similar if not the same concepts. In recent years, the term “lattice structure” has become common,

though additional terms such as “metamaterial”, “cellular structure”, or “mesoscale design” are

Page 262: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

233

also used, often interchangeably. In addition to these search terms regarding the lattice structure,

there are a variety of search terms for the manufacturing process including “additive

manufacturing”, “rapid prototyping”, “3D printing”, etc. Specific to this paper, the search was for

metal AM such as, “powder bed fusion”, “selective laser melting”, “direct metal laser sintering”,

and “electron beam melting”. To narrow the wealth of papers that these search terms can produce,

other search terms such as “mechanical performance”, “mechanical characterization”,

“mechanical properties”, etc. were included. No restrictions were placed on publication year of

the source. While the intent was to gather all published literature articles containing mechanical

characterization of metal lattice structures made with AM, it became difficult to find all current

and relevant literature due to a wide variety of terminology included in today’s research articles.

In various combinations, the aforementioned keywords were searched using Google Scholar and

Compendex. Additional references were also found through snowballing, a technique for

expanding literature reviews through exploration of relevant sources cited within the compiled

literature findings.

From each paper, three main data types were considered for inclusion: (1) analytical

models, (2) FEA, and (3) experimental. These three data types were defined as follows.

Analytical data was defined as mechanical performance or properties derived from

mathematical equations based on structural shear/normal forces or bending moments. For

example, models that were grouped into the analytical category were often derived using beam

theory, considering each individual strut within the unit cell topology. The result of the papers in

the analytical category was a continuous equation, generally relying on unit cell size and volume

fraction as inputs or independent variables and the predicted mechanical property being the

dependent variable.

Papers grouped in the FEA category were defined in terms of mechanical performance or

properties gathered through modeling and simulation of individual or arrayed unit cells. This data

Page 263: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

234

was available in research articles as either data points or extracted using a power law regression

fit as described in Section 6.2.2. Key features of this data type were the requirement to have full

information regarding unit cell topology, volume fraction or strut diameter, and material. To

normalize all results, raw data was divided by the corresponding bulk, or solid, mechanical

property. In cases where the reported data was not normalized, or where the bulk mechanical

property was not given in the source, the raw data was normalized using the mechanical property

values shown in Table 6.1.

Finally, papers in the experimental category were defined in terms of mechanical

performance or properties determined based on experimental testing of metal AM-fabricated

lattice structures. Experimental data was available in published literature as data points in tables

and on graphs. In some sources these data points represented an average from multiple tests;

however, in other sources, data points for all replicates in the experiment were reported as

opposed to a single average value. Similar to the FEA category, a clear description of the

geometry, volume fraction, and, where possible, as-built mechanical properties of solid material

specimens were gathered. In some cases, as-built strut diameter or volume fraction were reported;

however, nominal, or as-designed, values were also collected for consistency. As with the FEA

data type, if the reported data was not normalized, then the mechanical properties from Table 6.1

were used to normalize the data.

Table 6.1 For sources without bulk mechanical properties of the as-built material, reported within

the source, these generic AM mechanical property values were used to normalize data to create a

material independent comparison of unit cell topology.

AM Material Young’s Modulus (GPa) Yield Stress (MPa)

Co-Cr [371] 200 1060

AlSi10Mg [371] 75 270

Ti64 [379] 129 1007

SS 316L [371] 185 530

Hastelloy X [371] 195 630

Page 264: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

235

After gathering 117 research articles from the literature search, the total number of

articles was reduced to 69 using the following criteria:

• Paper must contain mechanical properties for quasi-static compressive behavior

• Paper must contain a clear depiction of unit cell topology through figures,

pictures, and/or textual description

• Paper must contain a clear description of unit cell volume fraction (or porosity)

• Paper must contain data for a standard/common unit cell topology (available in at

least one additional source)

A detailed breakdown of how the research articles were reduced is as follows. The total

number of articles was reduced from 117 to 109 by first removing those which were duplicates,

or reported data repeated in other sources. Next the total number of articles was reduced to 92 by

considering only those containing lattice structure mechanical performance and properties for

quasi-static compressive behavior. Though a few instances of lattice structure test data for tensile,

shear, bending, or combined loading conditions were found in the literature, the data was

extremely limited, and as such, only quasi-static compressive characterization data was included

in this work [213,217,323,380,381]. For other types of loading conditions, information regarding

lattice structure fatigue performance or high-strain deformation was recently becoming available

(c.f., [188,189,382–385]) but was not included in this work. The selection of papers was further

reduced to 69 due to insufficient information such as an ambiguous depiction of the unit cell

topology or volume fraction and non-standard unit cell topologies for which little data was

available [190,386,387]. Similarly, in some cases, part of the data was excluded based on the

limited data available on a unique topic, such as how the aspect ratio or rotation of the unit cell

topology affects mechanical properties [201,216,219,388]. In general, data was collected that

Page 265: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

236

considered regularly shaped unit cells, often cubic, that could be scaled by adjusting the unit cell

size or densified by increasing the thickness of struts or walls.

After the literature search and filtering, metal AM lattice structure data was obtained

from a total of 69 papers that were published between 1997-2019. Table 6.2 shows the references

and a summary of the unit cell topology, material, and the data type gathered from the paper. In

the table, the properties were abbreviated as normalized compressive effective modulus (EC),

yield stress (YS), Poisson’s ratio (PR), plateau stress (PS), and buckling stress (BS). Each of

these properties are defined in the context of this work in the results section (see Section 6.3.1).

The data types were abbreviated as analytical (A), FEA (F), and experimental (E).

Table 6.2 Summarized list of all sources for data compiled in LUCIE. In this table the unit cell

topology, properties recorded, and data type available are sorted by source.

Author Unit Cell Topology Property Data Type Material

Ahmadi [209] Diamond EC, YS, PR,

BS, PS A, E, F Ti-6Al-4V

Ahmadi [211]

Cubic, Diamond, Truncated Cube, Truncated Cuboctahedron, Rhombic

Dodecahedron, Rhombicuboctahedron

EC, YS, PS E Ti-6Al-4V

Arabnejad [364] Octet-truss EC, YS E Ti-6Al-4V

Ataee [389] TPMS Gyroid EC, YS, PS E CP-Ti

Beyer [390] BCC YS E AlSi10Mg

Bobbert [391] TPMS Gyroid, TPMS Diamond EC, YS E Ti-6Al-4V

Borleffs [392] Cubic, Diamond, Rhombic Dodecahedron, Truncated

Octahedron EC, PR A, F Ti-6Al-4V

Campanelli [393] FCCZ YS, PS E Ti-6Al-4V

Cheng [212] Rhombic Dodecahedron EC, YS E Ti-6Al-4V

Choy [388] Cubic EC, PS E Ti-6Al-4V

de Formanoir [366]

Octet-truss EC E Ti-6AL-4V

Deshpande [201] Octet-truss EC, YS A n/a

Gümrük [394] BCC EC, YS E, F SS 316L

Gümrük [380] BCC, BCCZ, F2BCC EC, YS E SS 316L

Page 266: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

237

Han [208] Cubic, BCC EC, YS E, F Co-Cr

Harrysson [395] Rhombic Dodecahedron EC, YS E, F Ti-6Al-4V

Hasib [396] Octahedron, Rhombic

Dodecahedron EC, YS E Ti-6AL-4V

He [369] Octet-truss EC, YS A n/a

Hedayati [381] Cubic, Truncated Octahedron,

Rhombic Dodecahedron, Diamond EC, YS, PR F Ti-6Al-4V

Hedayati [204] Rhombicuboctahedron EC, YS, PR A, E, F Ti-6Al-4V

Hedayati [203] Truncated Cube EC, YS, PR,

BS A, E, F Ti-6Al-4V

Hedayati [202] Cubic, Rhombic Dodecahedron,

Diamond, Truncated Octahedron EC, YS, PR A, F Ti-6Al-4V

Hedayati [206] Truncated Cuboctahedron EC, YS, PR A, F Ti-6Al-4V

Hedayati [205] Octahedron EC, PR A, F Ti-6Al-4V

Heinl [397] Diamond EC, YS E Ti-6Al-4V

Heinl [398] Diamond EC, YS E Ti-6Al-4V

Helou [192] TPMS Gyroid, TPMS Diamond EC E, F SS GP1

Hussein [220] TPMS Gyroid, TPMS Diamond EC, YS E Ti-6Al-4V,

AlSi10Mg, SS 316L

Leary [399] BCC, BCCZ, FCC, FCCZ, F2BCC EC, YS E AlSi12Mg

Leary [400] BCC, BCCZ, FCC, FCCZ EC, YS E Inconel 625

Lei [207] BCC, BCCZ EC, YS E, F AlSi10Mg

Li [401] BCCZ EC A AlSi10Mg

Li [402] Cubic, Rhombic Dodecahedron, G7 EC, PS E Ti-6Al-4V

Liu [403] Rhombic Dodecahedron EC E Ti2448

Liu [404] Rhombic Dodecahedron EC E Ti2448

Mazur [405] BCC, FCC, F2BCCZ EC, YS E Ti-6Al-4V

McKown [385] BCC, BCCZ EC, YS, PR E SS 316L

Mines [406] BCC EC, YS E Ti-6Al-4V

Mullen [407] BCC YS E CP-Ti

Murr [408] Rhombic Dodecahedron EC E Co-Cr

Murr [409] Cubic, Rhombic Dodecahedron, G7 EC E Ti-6Al-4V

Murr [410] Rhombic Dodecahedron EC E Inconel 625

Ozdemir [188] Diamond EC, YS E Ti-6Al-4V

Parthasarathy [411]

Cubic EC, YS E Ti-6Al-4V

Ptochos [412] BCC EC, PR A n/a

Rehme [413] FCC, FCCZ, BCC, BCCZ, F2BCC,

F2BCCZ YS E SS 316L

Sallica-Leva [414] Cubic EC, YS E Ti-6Al-4V

Shen [415] BCC, BCCZ YS E SS 316L

Smith [214] BCC, BCCZ EC, YS, PS E, F SS 316L

Page 267: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

238

Suard [365] Octet-truss EC E, F Ti-6Al-4V

Tancogne-Dejean [191]

Octet-truss EC, PR F SS 316L

Tsopanos [416] BCC EC, YS E SS 316L

Ushijima [215] BCC EC, PR A, F SS 316L

Ushijima [216] BCC EC, PR, YS A SS 316L

Van Grunsven [417]

Diamond EC, YS E Ti-6Al-4V

Wauthle [418] Diamond EC, YS E Ti-6Al-4V

Xiao [419] Rhombic Dodecahedron EC, YS, PS E Ti-6Al-4V

Yan [221] TPMS Gyroid YS E AlSi10Mg

Yan [218] TPMS Diamond EC, YS E AlSi10Mg

Yan [222] TPMS Gyroid EC, YS E SS 316L

Yan [223] TPMS Gyroid EC, YS E SS 316L

Yan [420] TPMS Gyroid, TPMS Diamond EC, YS E Ti-6Al-4V

Yanez [421] TPMS Gyroid EC, YS E, F Ti-6Al-4V

Yang [422] TPMS Gyroid EC E, F Ti-6Al-4V

Zadpoor [178]

Cubic, Diamond, Truncated Cube, Truncated Cuboctahedron, Rhombic

Dodecahedron, Rhombicuboctahedron, Truncated Octahedron, Octahedron, BCC, FCC

EC, PR, YS, BS

A n/a

Zaharin [423] Cubic, TPMS Gyroid EC, YS E, F Ti-6Al-4V

Zhao [424] Cubic, Rhombic Dodecahedron, G7 EC, PS E Ti-6Al-4V

Zhao [425] BCC EC, YS E Ti-6Al-4V

Zhu [426] Truncated Octahedron PR A n/a

6.2.2 Data Collection

Data was collected from the 69 papers that provided all of the required mechanical

property information for lattice structure characterization. The volume fraction of the tested unit

cell topology was required to compare data from different sources. Additionally, for experimental

and FEA data, the material needed to be given to normalize all the data. As noted earlier, data

was normalized by dividing the reported property value of the lattice structure by the bulk

Page 268: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

239

mechanical property for the as-built material. Normalization allowed comparison of data taken

from sources using different materials.

Data collection depended on how each paper presented their findings. If data was

explicitly reported as numerical values, then those values were simply catalogued along with the

other required information. In some cases, numerical data was not provided, and plotted graphs

were given instead. To extract the relevant data points, WebPlotDigitizer [427] was used to find

the approximate values of the plotted data. WebPlotDigitizer is a web-based image analysis tool

specifically designed to extract data from images of graphs and plots. The accuracy of the

extracted data was dependent on the figure image quality within the cited source and on the user

error in precisely selecting data points from the extracted image. WebPlotDigitizer has been

recognized as a viable method for data collection when the raw data is unavailable. Several

studies have been performed to verify and validate the use of such digital image data collection

tools [428–430]. Though digital image extraction of data points introduces uncertainty, one study

[429] reported that WebPlotDigitizer resulted in a Lin’s concordance value of 0.982, for which

0.99-1 is considered a near-perfect match [431]. From these studies, the extracted data was found

to be reliable, valid, have little absolute difference from the true values, and have excellent

consistency.

Meanwhile, in eight FEA data type sources [202–206,365,381,394], instead of reporting

specific simulation data points, the data was reported as points connected by line segments, i.e., a

segmented curve. Though in some cases a slight and sudden change in slope indicated the

likelihood of a data point, it was not possible to identify each of these data points on all plots.

Because individual points of the FEA simulation could not be distinguished, a power law

regression fit was used to match the segmented curve as best as possible and extract the data from

the graph. The regression fit of these segmented curves introduced some additional error;

Page 269: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

240

however, the primary intent of LUCIE was not to provide an exact mechanical property value, but

to provide property trends based on the unit cell type.

The power law regression was based on a popular property-volume fraction model for

cellular structures by Gibson and Ashby [136], where 𝜓 is the property, 𝑉𝑓 is the volume fraction,

and 𝛽1, 𝛽2 are constants determined by the fit of the model to the experimental data; see Eqn.

(3.13).

𝜓 = 𝛽1(𝑉𝑓)𝛽2

(6.1)

A modified power regression equation, with an additional constant, 𝛽3, was used

specifically for the Poisson’s ratio property, ν, to match data that did not intersect the origin (see

Eqn. (6.2)). For these eight sources with segmented curves, instead of individual data points,

WebPlotDigitizer was first used to extract data points along the curve. Second, the power law

regression fit coefficients for the extracted points were solved using Desmos [432]. Desmos is an

online graphing calculator that can fit a regression model to a set of data points.

𝜈 = 𝛽1(𝑉𝑓)𝛽2 + 𝛽3 (6.2)

6.2.3 Assumptions for Comparing Data

To compare data from different papers, many assumptions had to be made. Data was

collected only from papers that tested metal prints made with laser or electron beam powder bed

fusion processes. It was assumed that all types of metallic materials and the printing process

parameters were valid. We also assumed that data from different metallic powder feedstock was

comparable when normalized (e.g., different suppliers provide equivalent grades of Ti-6Al-4V

feedstock). Where possible, the normalized value was calculated using the mechanical property of

Page 270: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

241

the as-built material reported in the paper. In cases where the as-built material was not

experimentally characterized or given in the paper, bulk values were chosen from published data

sheets. The mechanical property values of different materials built with AM are given in Table

6.1.

Additionally, we assumed that all methods of testing, both experimental and FEA, as well

as analytical derivations can be compared. For FEA and analytical data, it was assumed that unit

cell size does not influence mechanical properties. This may seem like a limiting assumption, but

in many cases, papers did not report unit cell size for analytical models or FEA simulations. For

experimental data, unit cell size was recorded if it was given in the source. For analytical data

specifically, it was assumed that the derivations for volume fraction consider strut overlap and

were accurate. For many papers, data was extracted from plots, and it was assumed that data

taken from plots were valid within the resolution available. Though some papers measured the

actual volume fraction, or relative density, of the AM lattice structures, the theoretical volume

fraction, or as-designed value, was used in this work because it was reported more consistently in

literature than the as-printed volume fraction.

The potential sources of error include: experimental error within the reported results, low

image quality in figures used to extract data points with WebPlotDigitizer, human error in

precisely selecting data points in an image when gathering data, and the regression fits of the

segmented curves for some of the FEA data type data. While there were several potential sources

of error, the total error was assumed to be minimal when combining data from so many different

models and experiments across the literature. With all the variations in data due to testing

methods, AM process parameters, and materials, the purpose of the assumptions was to isolate

the effect of unit cell topology in mechanical property trends.

Page 271: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

242

6.2.4 LUCIE

The Lattice Unit-cell Characterization Interface for Engineering (LUCIE) was developed

in MATALB [344] as a design tool to compare and visualize mechanical properties of lattice

structures based on various unit cell topologies. LUCIE presents a large plot area along with

several adjustable settings to compare the lattice structure data that was collected from the

literature. The data can be sorted by the user adjusting the range of volume fractions and unit cell

sizes that are plotted. Few data points were available for a single volume fraction; so, a range

allows for more data to be viewed. Drop-down menus labeled “X Property” and “Y Property”

specify which lattice structure properties of interest are to be plotted on the X and Y axis,

respectively. Check boxes can then be used to select the type of data (Analytical, FEA, and/or

Experimental) and the unit cell topologies that are plotted. Multiple data types and unit cell

topologies can be selected to aid in comparison.

The FEA and experimental data collected from the literature search were stored in

spreadsheets. Numerical data catalogued the data’s source, unit cell topology, mechanical

property, volume fraction, unit cell size, numerical property value, units, bulk mechanical

property value, and the normalized property value. Regression data cataloged the data source, unit

cell topology, mechanical property, volume fraction min/max, unit cell size min/max, and

constants (𝑎, 𝑏, 𝑐) for the power regression equation; see Eqn. (6.2).

Equations for the analytical data were stored explicitly in a MATLAB [344] function. It

was typical for these equations to have either the volume fraction or aspect ratio (ratio of strut

radius to strut length) as the independent variable. In many cases, algebraic manipulations of the

analytical equations were required to use volume fraction or aspect ratio as the independent

variable.

Page 272: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

243

An example of inputs and outputs for LUCIE is shown in Figure 6.2. To use LUCIE, a

user must first set the desired data range values and unit cell topologies to be displayed along

with data types. Volume fraction and unit cell size are set either with the sliders or the text edit

boxes. For both volume fraction, a minimum and maximum value need to be specified so that

data can be selected from a range. For unit cell size, either “All” data can be plotted or a range

may be selected. For sources that did not directly specify the unit cell size of the test data, the

data points are included when “All” unit cell size data is selected. However, if a unit cell size

range is specified, only data with a known unit cell size is plotted. The desired properties for the

X and Y axes need to be specified along with the types of data and unit cell topologies are to be

plotted. Once all settings have been chosen, the MATLAB code filters all of the data so that only

data that satisfies the user’s specifications are plotted.

Page 273: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

244

Figure 6.2 Schematic showing the inputs and outputs for LUCIE

In the case where one of the properties selected was volume fraction, the data could be

plotted directly, because all the data types were reported, or calculated, as a function of volume

fraction. In the case where volume fraction was not selected as one of the properties to be plotted,

the properties were calculated by the specified range of volume fraction and unit cell size and

then parametrically represented.

Page 274: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

245

In the case of the analytical data type, data was represented by continuous functions. For

each specified lattice structure, the two functions that pertain to the selected properties of interest

were plotted parametrically with volume fraction as the input parameter.

In the case of the experimental and FEA data types, the data was represented by discrete

points, which were first combined into bins by volume fraction. Each bin contained a range of 5%

volume fraction, combining any data pertaining to that range. After the data from each property

type was sorted into bins, the data could be represented parametrically for each property by

volume fraction. In these cases, the data was plotted using two error bars that show the

maximums and minimums of the data in that bin. The location where two error bars intersect

denoted the average value. In the case of only a single data point within the bin range error bars

were not shown.

6.3 Results and Case Studies

6.3.1 Results of Literature Compilation

A summary of the type of data and properties for each unit cell topology is shown in

Table 6.3, and the unit cell topologies are illustrated in Table 6.4. Compiling all the data from 69

papers, 1,439 experimental data points were collected across 18 different unit cell topologies. For

the experimental data points, 41% were collected from tabulated values, and the remaining 59%

were extracted from digital images in the original sources using WebPlotDigitizer. In addition,

209 data points for finite element simulations were also collected. The majority of data was

available for the effective modulus (compressive) and yield stress of lattice structures; however,

there was also some data available in literature for Poisson’s ratio, buckling stress, and plateau

stress.

Page 275: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

246

Table 6.3 For each unit cell topology included in LUCIE, there are 6 different properties that can be

compared. The data type available for each unit cell topology are denoted by analytical (A),

experimental (E), or FEA (F). A dash (-) signifies that no reference was found that provided this

information for any of the three data types.

Unit Cell Topology Volume Fraction

EC YS PR BS PS

Cubic A A, E, F A, E, F A, F A E

Diamond A A, E, F A, E, F A, F A A, E

Truncated Cube A A, E, F A, E, F A, F A E

Truncated Cuboctahedron A A, E, F E, F A, F A E

Rhombic Dodecahedron A A, E, F E, F A, F - E

Rhombicuboctahedron A A, E, F A, E, F A, F - E

Truncated Octahedron A A, F F A, F - -

Octahedron A A, E, F E, F A, F - -

Octet-truss A A, E, F A, E F - -

BCC A A, E, F A, E, F A, E, F - E, F

BCCZ A A, E, F E, F E - E, F

FCC A A, E E A - -

FCCZ A E E - - E

F2BCC A E E - - -

F2BCCZ A E E - - -

TPMS Gyroid - E, F E, F - - E

TPMS Diamond - E, F E - - -

G7 - E - - - E

Page 276: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

247

Table 6.4 Name and associated picture for each lattice structure included in LUCIE

Unit Cell Topology Image Unit Cell Topology Image

Cubic

BCC

Diamond

BCCZ

Truncated Cube

FCC

Truncated Cuboctahedron

FCCZ

Rhombic Dodecahedron

F2BCC

Rhombicuboctahedron

F2BCCZ

Truncated Octahedron

TPMS Gyroid

Octahedron

TPMS Diamond

Octet-truss

G7

Page 277: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

248

As shown in Table 6.3, the mechanical properties for compiled data include normalized

compressive effective modulus (EC), yield stress (YS), Poisson’s ratio (PS), buckling stress (BS),

and plateau stress (PS). In Table 6.3, a dash (-) signifies that no reference was found that provided

this information for any of the three data types. One limitation of this work, as mentioned

previously, was the challenge with inconsistent terminology and methods for computing and

reporting mechanical properties. As such, a brief description of each property as we have defined

it follows.

Compressive effective modulus was defined as the slope of the linear portion of the

stress-strain curve during the initial loading or the linear portion of the unloading curve during

testing, although Ashby et al. [137] suggest that unloading effective modulus may be a more

accurate representation of the structure’s performance than the loading effective modulus. Yield

stress was defined as the first peak stress for stretch-dominated lattice structures, and was

generally the same as the plateau stress for bend-dominated lattice structures [250,400,405]. In a

few cases, sources may have instead reported yield stress based on a strain offset as opposed to a

first peak or plateau stress. Plateau stress was defined as the average stress value during the

cyclical collapse of the bending-dominated lattice structures or during the relatively level stress

that occurs in stretch-dominated structures [400,405]. Buckling stress was defined only in the

analytical data type because, in general, changes in slope on the stress-strain curve due to yielding

or buckling are difficult to distinguish. As such, the buckling stress of lattice structures was not

reported experimentally. The distinction or classification of stretch- or bend-dominated unit cells

was not categorized or recorded within LUCIE, but it was useful to understand what the yield or

plateau stress represents in different unit cell topologies.

Figure 6.3 illustrates each property and the differences between loading curves for

stretch-dominated and bending-dominated truss-like lattice structures. Bend-dominated structures

tend to have a peak stress at yield, and then the stress decreases as layers begin to collapse.

Page 278: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

249

Stretch-dominated structures tend to maintain a plateau in their stress strain curve as the struts

yield in tension and collapse prior to densification. Stretch-dominated and bend-dominated lattice

structures are determined based on their Maxwell number, which suggests how deformation

occurs in the structure [138,433].

Figure 6.3 Typical stress-strain curves for stretch-dominated or bend-dominated structures, labeled

with the relevant mechanical property terminology for the data compiled in this work

Regarding experimental data collected, the data may be grouped in four categories: (1)

“non-averaged”, (2) “non-averaged+uncertainty”, (3) “averaged”, and (4)

“averaged+uncertainty”.

Page 279: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

250

• “Non-averaged” data points were defined as data points that did not represent an

average from experimental replicates and did not include uncertainty reported in

the source. These data points were from sources that reported all of the data from

their experimental replicates as individual points as opposed to only the average

of the data, or from sources that did not have any experimental replicates. In the

LUCIE database for experimental data, “non-averaged” accounts for 65.8% of

the data points we collected.

• “Non-averaged+uncertainty” data points were defined as data points that did not

represent an average from experimental replicates but did include uncertainty

reported in the source. In a single source, replicates were not reported, but the

source estimated the uncertainty based on the testing equipment. In the LUCIE

database for experimental data, “non-averaged+uncertainty” accounts for only

0.7% of the data points we collected.

• “Averaged” data points were defined as data points that represent an average of

the experimental replicates but for which an uncertainty value was unavailable.

The unavailability of the uncertainty associated with the averaged data points

was for one of two reasons: (i) the uncertainty was not reported or (ii) it was not

visible for digital image extraction. For example, if uncertainty was reported as

error bars on a plot, then the error bars were occasionally hidden by the data

point marker they were attached to or other nearby markers, making it impossible

to accurately extract the uncertainty. In the LUCIE database for experimental

data, “averaged” accounts for 15.8% of the data points we collected.

• “Averaged+uncertainty” data points were defined as data points that represent an

average of the experimental replicates and included an uncertainty value. These

data points were often reported as tabulated values (e.g., 10 ± 1 MPa) or on plots

Page 280: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

251

that included error bars. In the LUCIE database for experimental data,

“averaged+uncertainty” accounts for 17.7% of the data points we collected.

From the collection of experimental uncertainty, 264 of 1439 (18.4%) data points were

reported with uncertainty. A histogram has been created to show the frequency of the relative

error for the data points reported with uncertainty; see Figure 6.4. The relative error was defined

as the uncertainty divided by the nominal value. For example, for a data point 10 ± 1, the relative

error was ± 10%. In cases where the error bars extracted from digital images were different on the

high or low side, the conservative, or larger, error value was taken. From the histogram of the

relative error, we find that when the experimental uncertainty was reported, 75% of the data had

an uncertainty of less than 10%. As there was only a select portion of the data, 18.4%, for which

experimental uncertainty was quantified in the literature, the uncertainty value was not directly

accounted for within the LUCIE database. As described in Section 6.2.4, error bars shown on a

plot within LUCIE denote the maximum and minimum values averaged from the bin of data for

that point. As more data becomes available with quantified uncertainty, future revisions allow the

user to restrict data to a specific cutoff threshold.

Page 281: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

252

Figure 6.4 Relative frequency of uncertainty in experimental data points. Uncertainty was

normalized by the nominal value to obtain the relative error. 75% of the reported uncertainty values

had a relative error of less than 10%.

6.3.2 LUCIE Case Study 1

Case Study 1 was a demonstration of the output from LUCIE. Figure 6.5 shows a

comparison of volume fraction and the normalized effective modulus using analytical,

experimental, and FEA data. The unit cell topologies plotted here are the BCC, Diamond, and

Truncated Cube for a volume fraction range and unit cell size of 0-50% and 0-10mm,

respectively, to capture a large range of data. It can be seen that for the analytical Diamond cell

data (red), three curves are plotted to reflect analytical results from different sources. In general,

these analytical curves represent the experimental and FEA data fairly well at low volume

fractions and at high volume fractions the range of data is large. In contrast, other analytical

models tend to underpredict or overpredict the experimental and FEA data. For example, in

Figure 6.5 the analytical BCC curve (blue) underpredicts the experimental data beyond a volume

fraction of 0.15. However, the analytical curve for the Truncated Cube overpredicts the

Page 282: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

253

experimental data reported for that unit cell topology and diverges from the FEA simulation

results.

Figure 6.5 LUCIE Case Study 1 output comparing analytical, experimental, and FEA data for the

BCC, Diamond, and Truncated Cube unit cell topologies. From the output you can visualize

differences in analytical models for a single unit cell topology, such as the Diamond, as well as gain

an understanding of the normalized effective modulus for different unit cell topologies.

6.3.3 LUCIE Case Study 2

Case Study 2 contains only experimental data from LUCIE; see Figure 6.6. This figure

plots the volume fraction and normalized yield stress for Diamond, Truncated Cube, Truncated

Cuboctahedron, Rhombic Dodecahedron, Rhombicuboctahedron, BCCZ, and TPMS Gyroid unit

cell topologies. The ranges of volume fraction and unit cell size were set to 5-40% and 1-5mm,

respectively. The plot indicates a general area where a particular unit cell topology’s data exists

and how it compares to other unit cell topologies and types of data. For example, the Diamond

unit cell topology was reported to have a lower normalized yield stress than the Truncated Cube

Page 283: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

254

or TPMS Gyroid; however, it was on par with reported data for the Rhombic Dodecahedron. By

making these comparisons, a user can easily update the settings in the interface to narrow the

search for a particular unit cell topology.

Figure 6.6 LUCIE Case Study 2 output shows a comparison of 7 different unit cell topologies and the

variability that exists in experimental data. The error bars around a single point represent a small

“bin” of data. The point itself represents the average value within the bin, while the error bars

denote the maximum and minimum reported values for each property.

LUCIE can benefit designers by allowing them to easily compare unit cell topologies. For

example, based on the plot shown in Figure 6.6, a designer who was previously considering a

diamond unit cell topology would see that by switching to a TPMS gyroid unit cell topology there

are opportunities for further reducing weight or increasing the normalized yield stress value.

Specifically, starting from a diamond unit cell topology with a volume fraction of 0.35-0.4, a

designer can achieve the same normalized yield stress with nearly 50% weight savings by

switching to a TPMS gyroid unit cell topology with a volume fraction of 0.2. Similarly, by

Page 284: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

255

switching from the previously mentioned diamond unit cell to a TPMS gyroid unit cell topology

with volume fraction between 0.35-0.4, the normalized yield stress value is increased by 76%

with no change in weight.

6.3.4 LUCIE Case Study 3

Case Study 3 illustrates how to use LUCIE as a design tool to aid unit cell selection by

providing information to differentiate unit cell topologies and their performance. Consider the

design requirements for a lattice structure with a target density, normalized effective modulus,

and normalized yield stress of 0.2-0.25, 0.02-0.03, and 0.05 or greater, respectively. LUCIE was

used to plot experimental data for volume fraction and normalized effective modulus of the 18

different available unit cell topologies; see Figure 6.7. Next, the user locates the region of interest

on the plot according to the design constraints on volume fraction and normalized effective

modulus. The region of interest is highlighted in Figure 6.7 by a red box with a dashed line. From

this plot we can identify six potential unit cell topologies, whose average normalized effective

modulus falls within the region of interest. This reduces our selection from 18 to 6 unit cell

topologies: Cubic, BCC, Truncated Cuboctahedron, and Rhombicuboctahedron, TPMS Gyroid,

and Octet-truss.

Page 285: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

256

Figure 6.7 LUCIE Case Study 3 output from LUCIE for comparison of the lattice structure

properties of different unit cell topologies. The red box with a dashed line highlights the region of

interest (EC 0.02-0.03 and Volume Fraction 0.2-0.25) for determining appropriate unit cell selection.

A new plot was generated by selecting only the six unit cell topologies identified from

Figure 6.7, with a comparison of normalized effective modulus and normalized yield stress with a

volume fraction restriction of 0.2-0.25; see Figure 6.8. It can be seen that there was a clearer

distinction between the possible cells. The Cubic cell has the highest average reported normalized

yield stress followed by the TPMS Gyroid, Rhombicuboctahedron, Truncated Cuboctahedron,

and finally the BCC. No normalized yield stress data was available in this range for the Octet-

truss. With a stipulation that the normalized yield stress be greater than 0.05, the designer could

move forward with either a Cubic, TPMS Gyroid, or Rhombicuboctahedron. One consideration

could be the large range of normalized yield stress that has been reported for the TPMS Gyroid,

directing a user to further explore how the normalized yield stress can be high or low for similar

volume fractions. Further distinction can be considered based on desired print tolerances, unit-cell

pore sizes, etc. For example, the Rhombicuboctahedron has more struts than the Cubic unit cell,

Page 286: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

257

meaning that, for the same volume fraction and unit cell size, the strut diameters and unit cell

topology pore sizes are smaller than those of the Cubic cell. If a small unit cell topology pore size

is critical to the application, the Rhombicuboctahedron would likely be a better choice than the

Cubic unit cell topology; however, if minimum feature size is a concern, the Cubic unit cell

topology has larger strut sizes and pores than the Rhombicuboctahedron making it easier to print

remove powder. For an AM designer, rather than an expensive experimental trial with 18

different unit cell topologies, LUCIE was able to quickly narrow the search to only three unit cell

topologies that could be compared or prepared for further testing, resulting in valuable cost and

time savings.

Figure 6.8 LUCIE Case Study 3 plot comparison of normalized effective modulus and normalized

yield stress for comparison of the six unit cell topologies that were down selected from Figure 6.7.

The thumbnails of the unit cell topologies have been added to the plot; from top to bottom they are

Cubic, TPMS Gyroid, Rhombicuboctahedron, Truncated Cuboctahedron, BCC. No yield stress data

was available for the Octet-truss at this volume fraction range.

Page 287: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

258

6.4 Discussion

While the data presented in the interface can be used by designers to compare the lattice

structure properties to improve unit cell selection, there were a number of challenges that remain

and limitations that exist in the current approach. AM remains a continuously changing and

evolving process with a wide range of process parameters, materials, and build strategies still

being developed. This work compiled data by unit cell topology, but not by the material or

processing conditions. The intended result was not to say that all data for a unit cell topology was

processed with a single set of process parameters; rather, it shows the range of mechanical

properties that can be achieved with a single unit cell topology. Even the analytical models that

have been developed differ based on their underlying assumptions and the experimental data

differs even further; see the diamond cell analytical models shown in Figure 6.5 as an example.

In cited sources, several reasons for differences between analytical, finite element, and

experimental data were suggested. Analytical models tend to overestimate effective modulus

when compared to finite element and experimental data and deviate more as volume fraction

increases. Analytical models derived using the Euler-Bernoulli method ignore shear and

rotational inertia which decrease the effective modulus [189,203,204,206]. Shear and rotational

inertia become more influential as volume fraction increases where more unit cell volume exists

at strut intersections. Analytical and FEA values for yield stress can overlap if critical struts

experience no bending and therefore were unaffected by shear and rotational inertia [204]. Finite

element models using the Timoshenko method account for shear and rotational inertia.

As for experimental data compared to analytical and FEA data, it was stated in the

literature that imperfections in the printing process were responsible for the biggest discrepancies.

For example, surface roughness as a result of unmelted or semi-melted powder can cause

variations in strut radius and cause stress concentrations [203–206,208,392,394,395,423].

Page 288: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

259

Additionally, quality of print for vertical, inclined, and horizontal struts vary, and this was not

captured by finite element or analytical models, which assume perfectly cylindrical and

homogeneous struts [206]. Meanwhile, strut alignment often differs between experimental and

finite element models as well. Analytical and finite element models do not account for slight

misalignments among struts and loading conditions [192,392]. Other causes for disagreements

may be due to the number of unit cells tested (finite or infinite) or lack of accounting for strut

interference during densification [203,214].

Different approaches have been proposed to increase the accuracy of analytical and finite

element models. Several sources try to account for imperfections in their models by changing the

as-designed strut radius to match a geometric average of a physical radius or by applying a

randomly varying strut radius [365,392]. One study improved their finite element model by using

reconstructed struts based on CT-scans which had greater accuracy than as-designed and average

radii models [207]. Having misaligned struts in finite element simulations has also been proposed

to improve the model [392].

As another example of the variation that was present in the literature, consider the gyroid

data presented in Figure 6.6. The range of data that was available from 0.1-0.15 volume fraction

had a large range in the reported yield stress as shown by the large vertical error bars. This

finding suggests that a difference in testing conditions, build quality, definition of the gyroid, or

processing parameters had major effects on the yield stress of the gyroid unit cell topology. The

variation seen in the gyroid could lead to several outcomes. First, one researcher could see the

data and take it as an argument for standardization of process parameters and testing procedures.

While standardization is critical, by combining all the results into a single interface, the range of

capabilities of the gyroid are represented. A second researcher could take the result and wish to

understand how the gyroid can be tailored to meet a specific yield stress criterion. By reviewing

Page 289: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

260

the original literature sources compiled in LUCIE, the ability for researchers to explore the

capabilities of the gyroid lattice structure is simplified.

Another factor in the variability of data could be the variations in testing conditions such

as the number of unit cells included in the array, boundary conditions in the compression test,

strain rate, etc. For example, ISO 13314:2011, the standard that addresses mechanical testing of

porous and cellular metal structures, recommends a periodicity of 10x in each direction to

minimize the effects from boundary conditions. However, as mechanical testing of lattice

structures has not been formalized, many authors do not strictly follow this guideline and test

samples with fewer than ten unit cells in any of the three directions. In this work, test samples

with fewer than ten unit cells in each direction were not excluded, resulting in data that may be

skewed by differences in the impact of boundary conditions on testing results. For additional

detail on the impact of boundary conditions on testing, see Section 5.2.3.3 and Section 5.5. As

with standard materials and testing, consistent standards applicable to lattice structures are needed

to ensure that the scientific community—and eventually designers and engineers—can regularize

the testing methods to achieve consistent results.

Another inherent challenge with AM is the design freedom that is available. Because

there is so much design freedom, there are infinite minor variations on a single unit cell topology,

such as rounding or fillets at node locations, making it impossible to name and track each

individual unit cell topology. The purpose of this work was not to say that the optimal unit cell

topology is found within the 18 topologies included in LUCIE. In our opinion, there is a large gap

between what can be designed and quantified for analysis and what can be printed in metal using

current laser and electron beam powder bed fusion technology. AM processes can print

complexity that designers would have never dreamed of or imagined previously. However, for the

designer to feasibly characterize such infinite possibilities and complexity into a functional and

reliable component is currently not possible. The intention of LUCIE was to help bridge this gap

Page 290: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

261

by providing a database to allow researchers and designers an understanding of regular unit cells

that have been characterized through analytical, experimental, and finite element analysis. In this

work, we have grouped similar unit cell topologies where possible, though minor variations in

geometric models do exist. As such, the purpose of LUCIE was visualization of performance

trends based on the unit cell type.

Related to the challenge imposed by unit cell topology design freedom, comes the

challenge of naming convention. Several authors have tried to address this issue for lattice

structures and unit cell topologies and it remains an ongoing battle [198–200]. When compiling

our data, the unit cell topology naming conventions were similar; however, variations were seen

across different research articles and software libraries. In this work, we have attempted to use the

most common name found in literature for describing the unit cell topology.

The amount of data available is increasing all the time but remains limited, especially for

some unit cell topologies. In the unit cell selection example, Case Study 3, represented by Figure

6.7 and Figure 6.8, there was lattice structure data available that fit the design requirements;

however, in some cases data may not be available because there are many gaps in the published

literature. For example, the Octet-truss was initially identified as a potential candidate in Case

Study 3, but no data was available for yield stress in the specified volume fraction range. The

focus of this initial work was common unit cell topologies, or those for which more than a single

source could be found. As a result, less common and modified versions of common lattice

structures were excluded for lack of sufficient data. Figure 6.9 shows the number of sources used

for each unit cell topology. With the amount of published data increasing all the time, we expect

to be able to continue expanding the database for LUCIE and the research community’s

understanding of the range of mechanical properties that may be achieved with AM lattice

structures.

Page 291: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

262

Figure 6.9 Number of sources found for each unit cell topology. The results show the amount of

lattice structure data available is limited for many common unit cell topologies.

As the focus of this work was identifying the material independent mechanical property

trends of lattices structures based on the unit cell topology, all materials were combined through

normalization. At present, sorting by material is not available to a front-end user due to the

limited amount of data available for each material. With continued expansion of the database for

LUCIE, future work may include sorting or categorizing data by material as an option to further

refine the results in LUCIE.

Another limitation of LUCIE is that the exact testing conditions (testing machine type,

crosshead rate, boundary conditions, etc.) and AM processing parameters are not stored within

the current LUCIE database or available to the user on the front-end. For the testing conditions,

the information is available in the cited sources. In contrast, in terms of fabrication, few papers

report the specific details of the AM process parameters, often due to the proprietary information

of the AM machine manufacturer. Of those papers that do report AM process parameters, the

Page 292: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

263

amount and type of information that was reported is inconsistent across different sources. The

ability to incorporate testing conditions and AM process parameters for each data point in LUCIE

in a useful manner was thus limited because each data point represents numerous papers, each

with their own testing methods and AM process parameters. Future work for LUCIE could also

address how to make the sources for each combined data point accessible to the user, along with

information regarding testing methods and AM process parameters for comparison of how these

factors affect the mechanical properties.

Based on the research articles which have been compiled into the LUCIE database, a

number of future directions have been identified for short-term and long-term lattice structure

research. In the short-term, testing methods, naming conventions, and common practices should

be developed and standardized to allow researchers to collaborate easily and compare

experimental data and analytical and simulation results. The issue of lack of standardization

makes it very difficult for researchers to understand the current state when published works are

obfuscated by unique naming conventions. Standardization across naming conventions and

testing appears to be the low hanging fruit that can help accelerate lattice structure research for

AM. In the long-term, as the interest in lattice structures grows, research initiatives should

include exploration of:

1) additional experimental loading conditions such as tensile, shear, and fatigue,

2) novel modeling approaches to account for hierarchical complexity,

3) incorporation of manufacturing variability in modeling and simulation,

4) novel unit cell topologies through optimization,

5) post-processing opportunities and effects, and

6) industrial-proven case studies of lattice structure advantages.

Page 293: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

264

Additional methods to model and predict the effect of lattice structures within the full

component will be an enabling factor for driving forward industrial applications. As discussed

previously, the majority of the present literature was focused on compressive behavior, although

most components undergo a variety of loading conditions. Non-compressive behavior, especially

cyclic and fatigue loading, is a major concern among researchers for the applications of lattice

structures. Finally, learning how to further improve lattice structures with novel topologies or

post-processing methods continue to advance the opportunities available for lattice-enabled

components.

There are ongoing efforts to make LUCIE publicly available for all researchers and

engineers. Currently, LUCIE can be downloaded as a standalone application for installation on a

Windows operating system (see https://sites.psu.edu/edog/lucie/). This standalone version

requires the MATLAB Compiler Runtime which will be downloaded and installed for free

automatically during the installation process, meaning that a MATLAB license is not required to

use the application. In addition, ongoing efforts include development of a web-based version of

LUCIE that will allow researchers to utilize the application without having to download and

maintain an up-to-date version of the application. The intent of the web-based tool was to

simplify access to LUCIE with additional plans to allow researchers to be able to submit their

own data for incorporation into the database.

6.5 Closing remarks

Through review of over 69 papers, the largest known mechanical property database for

lattice structures fabricated with metal AM was compiled and the Lattice Unit-cell

Characterization Interface for Engineering (LUCIE) was developed to provide Ashby-style plots

for unit cell selection of AM lattice structures. Researchers and AM designers stand to benefit

Page 294: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

265

from simplified methods to observe and understand unit cell topology effects on lattice structure

mechanical properties through LUCIE and similar interfaces and databases. Though the amount

of data and lack of testing standards were limiting, ongoing efforts to characterize lattice

structures will continue to improve the models and experimental data available.

Page 295: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

266

Chapter 7 Conclusions and Future Work

In this dissertation, two systematic design optimization approaches were presented: a

shape matching approach for a compliant deployable surgical tool for radiofrequency ablation

(RFA) of tumors and the development of specialized lattice structures for additive manufacturing

(AM). This final chapter of the dissertation summarizes the work completed regarding each

objective. Next, major conclusions drawn from work are given. Following the major conclusions,

research contributions to the scientific community and broader impacts of the research are stated.

Finally, recommended future research directions are addressed.

7.1 Summary of Objectives and Tasks

7.1.1 Objective 1: Develop systematic optimization approach for compliant deployable

radiofrequency ablation electrodes

1.1 Design deployable electrode for endoscopic radiofrequency ablation

1.2 Develop finite element model to simulate radiofrequency ablation in soft tissue

1.3 Couple finite element with optimization approach for design of electrode based on

tumor geometry

1.4 Evaluate the deployable electrode design

1.4.1 Assess deployment of early prototypes in tissue phantom

1.4.2 Validate optimized electrode design through ex-vivo experimentation

To improve treatment for endoscopic RFA of tumors, a novel compliant deployable RFA

electrode was presented and optimized to match the treatment zone with the tumor geometry. The

deployable electrode design is based on an array of compliant tines which compress and stow

inside an endoscopic needle. The endoscopic needle serves as a sheath and then the electrode is

advanced out of the sheath and a circular array of tines deploy into the cancerous tissue. A finite

Page 296: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

267

element model which combines the electrical and thermal interactions of an electrode heating the

surrounding tissue has been developed in COMSOL Multiphysics version 5.2a [343]. The finite

element model predicts the ablation zone that develops around a deployed electrode geometry.

Coupling the finite element model with an optimization algorithm, the deployable electrode

design was optimized for treatment of a 2.5 cm spherical tumor. Comparing computational results

of a baseline electrode with 25% treatment efficiency, the deployable electrode design was shown

to increase treatment efficiency to 71-87%. Prototypes of the electrode were fabricated using AM

and laser micromachining. These prototypes were evaluated for deployment and later used for

validation in a thermochromic tissue phantom. Validation of the computational models showed a

strong dependence on temperature dependent properties that can be averaged over the simulation

time to improve computational efficiency. With corrections to the computational model to

account for tissue damage at elevated temperatures, the computational model was shown to

accurately replicate the ex-vivo testing results. At the time of prototyping, AM was not capable of

producing a functional deployable endoscopic electrode due to the minimum feature size and

material selection. However, the materials and processes available are constantly evolving and

this approach should be reconsidered for mass customization of electrodes possibly through AM

of Nitinol or metal binder jet processes. For additional details regarding Objective 1 and its

associated tasks, see Chapter 2.

7.1.2 Objective 2: Develop three-dimensional ground structure topology optimization for

unit cell DfAM

2.1 Develop three-dimensional ground structure topology optimization for unit cell DfAM

2.2 Establish systematic optimization method for unit cell generation

2.2.1 Generate set of potential objective functions and constraints

2.2.2 Quantify manufacturing considerations for metal AM

2.3 Validate optimization with benchmark problems

Page 297: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

268

With the increasing interest of AM for fabrication topology optimized structures, there is

a need for systematic approaches for the design of lattice structures. For non-stochastic lattice

structures, the most basic repeating pattern is the called the unit cell. In current practice, there is

little known about how the choice of unit cell topology affects the performance of a lattice

structure. Using commercial software, the choice of unit cell topology is typically based on trial

and error with little thought to how the choice will influence the structure. AM enables

customization and optimized lattice structures, but there is a lack of design optimization methods

that incorporate DfAM restrictions to ensure manufacturability.

Additive Lattice Topology Optimization (ALTO) is an application-agnostic systematic

approach for generating unit cell topologies. The foundation of this work relies on ground

structure topology optimization. Ground structures are particularly well suited to the design of

lattice structures because of their computational efficiency and they are a good representation of

traditional strut and node type unit cell topologies. The ground structure representation of the unit

cell has been developed in MATLAB and is paired with either a gradient-based or evolutionary

optimization algorithm. The gradient-based algorithm was utilized for 2D optimization problems

with two or fewer objectives. It was found to be very sensitive to the initial guess and the order of

magnitude of the objectives, requiring problem specific scale factors to equilibrate different

objective functions. The evolutionary algorithm was used for 3D optimization problems with

three or more objectives. It was found to be very robust to including different objective and

penalty functions for various types of optimization problems was able to successfully solve

optimization problems with more than one hundred design variables.

For the optimization, a library of objective, constraint, and penalty functions have been

developed. DfAM is incorporated throughout ALTO in various objective functions (i.e., unit cell

volume, minimization of relevant elements, minimization the powder removability factor),

constraint functions (i.e., minimum feature size), ground structure restrictions (i.e., build angle),

Page 298: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

269

and penalty functions (i.e., overlap, unsupported–S, unsupported–N). The DfAM considerations

reduce the need for design interpretation prior to printing the optimized solution. Specific

considerations to ensure manufacturability include bridging of horizontal elements, a no-build

range for the build angle, and minimum feature size for printing. Validation of the ground

structure formulation and optimization were performed using several traditional topology

optimization benchmark problems. Benchmark problems were solved for minimization of strain

energy, maximization of thermal conductance, and predicting macroscale properties from a

representative volume element through homogenization. In each benchmark, ALTO was able to

replicate results similar to the benchmarks from literature, validating its formulation for DfAM

cases studies. For additional detail regarding Objective 2 and its associated tasks, see Chapter 3.

7.1.3 Objective 3: Explore multi-functional lattices through case study on unit cell

generation for thermal conductance and minimum strain energy

3.1 Define constraints and process limitations

3.2 Generate optimal lattice structure with and without DfAM

3.3 Fabricate lattice structure with DfAM considerations

3.4 Evaluate solutions and revise ground structure topology optimization

Multi-objective design optimization offers a unique opportunity to enhance performance

through multi-functional lattices. Two case studies were presented for multi-functional lattices

that combine minimization of strain energy and maximization of thermal conductance. Using

various constraints and ground structure restrictions the case studies each produce designs that

exhibit the tradeoffs required to balance two conflicting objectives. Case Study 1 from Chapter 4

clearly illustrated the benefits of DfAM considerations. With DfAM, the optimized solution was

immediately manufacturable and was fabricated using two different AM processes. Without

DfAM, the optimized result would require significant design interpretation prior to fabrication.

Page 299: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

270

Case Study 2 from Chapter 4, demonstrated the Pareto-optimal solutions from which a designer

can select a design. Gradient-based algorithms were found to require significant time upfront to

determine appropriate weighting values and optimization parameters to balance the two

objectives. On the other hand, evolutionary algorithms were found to be robust in solving multi-

objective optimizations with less control over the development of the Pareto front. For additional

details regarding Objective 3 and it associated tasks, see Chapter 4.

7.1.4 Objective 4: Explore homogenized lattices through case study on generation of unit

cells with tailored compliance

4.1 Define constraints and process limitations

4.2 Generate optimal lattice structure with tailored properties

4.3 Fabricate and test to determine mechanical properties

4.4 Characterize fabricated samples with mechanical testing

4.4.1 Compare boundary conditions of homogenization with experimental

characterization

Lattice structures are of particular interest to the topology optimization community for

the design of materials with tailored compliance. Three case studies were presented for finding

structures that match a specific target constitutive matrix. Using a variety of optimization

objective, constraint, and penalty functions, these case studies add to the range of applications for

ALTO. Case Study 1 from Chapter 5 demonstrated a novel powder removability factor as an

objective function. The powder removability factor encouraged solutions with large spacing in

the unit cell topology to ease powder removal. After fabrication of the optimized and baseline

unit cell topologies, the optimized solution demonstrated improved powder removability

compared to the baseline unit cell through mass measurement and CT scan analysis. Experimental

compression testing was also used to validate the homogenization calculation within ALTO. With

corrections for the boundary conditions and differences between the as-designed and as-printed

Page 300: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

271

models, the homogenization predicted the effective modulus of two different unit cell topologies

to within 6%. Case Study 2 from Chapter 5 generated multiple optimized solutions, each for a

scaled version of the target constitutive matrix. The target volume fraction was constant,

requiring the optimization to rearrange, as oppose to simply remove, material. The result was

three different stiffness unit cells that all had the same weight by embedding mechanism-like

behavior into the unit cell topology. This enables lattice structures with a non-uniform stiffness

and uniform weight distribution. Fabricated samples were also experimentally tested and the

results further support boundary conditions corrections discussed with Case Study 1. Case Study

3 from Chapter 5 demonstrated generation of an orthotropic material with three times the

effective modulus in one direction as compared to the other. In addition, the DfAM penalty

function and relevant diameter limit were explored for their impact on the optimization problem.

For additional details regarding Objective 4 and it associated tasks, see Chapter 5.

7.1.5 Objective 5: Develop mechanical property database for metal lattice structures from

current literature containing analytical, finite element, and experimental data

5.1 Perform literature search for metal lattice structure data

5.2 Collect data from selected papers

5.3 Create user interface for generating mechanical property charts of collected data

While AM enables fabrication of optimized unit cell topologies, there is a need to be able

to characterize and understand standard unit cell topologies. Researchers around the world are

characterizing lattice structure through analytical, finite element, and experimental data.

Compiling data from 69 sources, the Lattice Unit-cell Characterization Interface for Engineers

enabled visualization of mechanical properties of metal lattice structure fabricated with AM. As a

result of the compilation, LUCIE contains data for 18 different unit cell topologies. The Ashby-

style material property charts are generated from over 1400 experimental data points, 200 FEA

Page 301: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

272

data points, and 45 analytical models. To the author’s knowledge, LUCIE is the largest publicly

available database of its kind. LUCIE can be downloaded as a standalone application for

installation on a Windows operating system (see https://sites.psu.edu/edog/lucie/). This

standalone version requires the MATLAB Compiler Runtime which will be downloaded and

installed automatically during the installation process; a MATLAB license is not required to use

the application. The interface enables researchers and scientists to understand the trends for

mechanical property data sorted by the data type, unit cell topology, unit cell size, and volume

fraction. It serves to help bridge the gap between formal optimization of the unit cell topology

with ALTO and the current practice of selecting blindly from a unit cell library. For additional

details regarding Objective 5 and it associated tasks, see Chapter 6.

7.2 Primary Conclusions

From the objectives and tasks in this dissertation, fifteen conclusions were drawn

regarding the work completed:

1. Deployable endoscopic electrodes are feasible for expanding RFA treatment zones and

have the potential to significantly impact the size of tumors that can be treated with a

single insertion of the electrode. The deployable electrode designs presented increased

treatment efficiency from 25% to 71-87%.

2. Temperature dependent properties such as electrical conductivity play a significant role

in the development of the ablation zone surrounding an electrode and must be addressed

in computational models.

3. The degree of deployment for a deployable electrode demonstrated no statistical

difference based on insertion speed, however, a minor trend appeared to show decreasing

deployment with increasing speed above 10 mm/s.

4. At the time of initial investigation, mass customization of endoscopic electrodes with L-

PBF was limited due to the minimum feature size of large-scale production machines and

material selection but is worth reconsidering as available material and processes continue

to improve.

Page 302: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

273

5. ALTOs unit cell topology optimization framework is capable of handling a wide variety

of applications as demonstrated through multi-functional lattices and various tailored

compliance lattice structures in the case studies presented.

6. The gradient-based optimization algorithm fmincon was capable of solving multi-

objective optimization problems with 2 objectives and 30 design variables.

7. The evolutionary algorithm Borg was capable of solving multi-objective optimization

problems with 4 objectives and 120 design variables.

8. The ground structure formulation and optimization of ALTO were validated through

common topology optimization benchmarks for minimization of compliance,

maximization of thermal conductance, and homogenization.

9. The novel powder removability factor was found to be a sufficient comparison metric for

manufacturability of a lattice structure.

10. Appropriate homogenization boundary conditions for the test strain cases applied to the

unit cell are strongly dependent on the number of layers and overall size of the intended

lattice structure.

11. AM-based correction factors for differences between as-printed and as-designed

structures, such as the BOC or SCC, can be incorporated directly into ALTO for

mechanical property prediction.

12. Through compression testing of printed lattice structures, the ground structure model and

homogenization method were validated experimentally.

13. DfAM optimization of unit cell topology simultaneously improves our ability to design

novel lattice structures and their manufacturability using AM.

14. Current experimental characterization of metal lattice structures fabricated with AM is

primarily limited to static compression testing without a standardized protocol for details

such as specimen size and strain rate.

15. Discrepancies between predicted mechanical properties from analytical and finite

element analysis as compared to experimental characterization remains a challenge for

advancing lattice structures to practical components.

7.3 Research Contributions

The research contributions of this work have been summarized into a list of seven key

points below. Following the list of research contributions, the intellectual merit for the scientific

community and broader impacts to the general community are addressed. With regards to the

Page 303: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

274

intellectual merit and broader impacts, the discussion is divided based on the two systematic

optimization methods developed: a shape matching approach for a compliant deployable surgical

tool for radiofrequency ablation of tumors and the development of specialized lattice structures

for additive manufacturing.

1. Validated systematic design approach for finite element-based optimization of

radiofrequency ablation electrodes

2. Developed an application-agnostic DfAM framework for lattice unit cell design and

optimization

3. Demonstrated first known DfAM powder removability optimization objective for lattice

structures

4. Demonstrated unit cell DfAM framework through multiple 2D and 3D examples

5. Generated novel unit cell topology with improved manufacturability and the same

orthogonal effective modulus properties as traditional unit cell

6. Developed build angle dependent staircase correction factor for the reduced effective

cross-sectional area of loading for unit cell struts

7. Established largest known database for analytical, finite element, and experimental

mechanical characterization of metal lattice structures into an interactive interface.

7.3.1 Electrode Optimization for Radiofrequency Ablation

Research of RFA has included numerous modeling methods for understanding how tissue

properties, power settings, and proximity to surrounding structures affect the treatment efficiency.

A few methods have also considered using these models for optimizing placement of the

electrode to improve treatment efficiency. While these models have improved our understanding

of RFA, there is a lack of systematic design approaches which take full advantage of these

models, optimization, and the increased manufacturing capabilities available.

For the scientific community, this work provides an experimentally validated novel

systematic design method for finite element-based optimization of the electrode geometry. The

Page 304: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

275

electrode geometry was optimized such that the ablation zone generated matched a target

geometry. In ablation technology research, the completed research fills the gap of using ablation

zone modeling as a design tool by providing validated models which predict ablation zones for

RFA electrodes as a methodology for optimizing them for a target shape. Though a very specific

case was considered as an example, EUS-RFA for a 2.5cm spherical tumor in pancreatic tissue,

the approach is applicable to tumors of any size/shape and to other locations where RFA may be

applied such as the lungs, liver, and kidneys. This systematic approach opens the doors for future

research into personalized medical care as it can easily be extended to customized non-spherical

and non-symmetric target shapes.

For a broader impact on the community, these models will benefit both doctors and

patients. This research moves surgeons one step closer to being able to provide truly customized

surgical treatment by enabling more advanced surgical tools. With optimized surgical electrodes,

surgeons can efficiently destroy cancerous tissue, preserving the surrounding healthy tissue. With

increased CT-imaging for medical diagnosis, many patients are learning of tumors in earlier

stages and seeking curative treatment options. While surgical resection is currently the most

effective option, a complete surgical resection of the tumor is often not possible due to

complications associated with the location of the tumor or the negative impact that an open

surgery would have on a patient in poor health. Improving less invasive treatment options than

open surgery through EUS-RFA, while simultaneously increasing efficiency, increases the odds

of survival for these rising numbers of cancer patients.

7.3.2 Additive Lattice Topology Optimization

With the advancing capabilities of AM, our ability to manufacture has exceeded our

current ability to design. AM processes enable access to the entire design envelope of a

Page 305: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

276

component and can fabricate designs that are difficult or impossible for traditional modeling tools

to analyze. While AM offers extensive freedom, there are additional design rules and process

limitations that have to be considered. There is a lack of systematic design tools and methods for

optimizing performance while accounting for AM process constraints, evident by numerous

research groups exploring topology optimization for AM, development of DfAM frameworks,

and a push from industry to be able to maximize performance of every component. In particular,

mesostructure design for lattice structures have broad applications; however, they are not well

understood or characterized and often used without consideration for why one lattice type is

chosen over another.

For the scientific community researching AM and DfAM, novel contributions of the

work completed are an application-agnostic systematic optimization method for unit cell topology

optimization, the powder removability factor (PRF) for improving manufacturability, and LUCIE,

the largest known database for mechanical characterization of lattice structures fabricated with

metal AM. The research combines ground structure topology optimization and DfAM to enable

generation of novel unit cell topologies with application-specific properties. Using an

evolutionary algorithm, the approach fills the gap of a robust methodology to incorporate a

variety of objective, constraint, and penalty functions that improve manufacturability for multi-

functional lattices. For example, in L-PBF, powder removal from lattice structures is a continuing

concern. The PRF developed as part of this work offers the first powder removal metric for strut

and node type lattice structures, assessing the manufacturability and post-processing of the lattice.

Generation of a lattice structure with the PRF as an objective has been shown to improve

manufacturability. For common lattice structure types, numerous research groups around the

world are generating much needed analytical, finite element, and experimental data for

characterization of lattice structures. LUCIE offers the largest compilation of this data in a format

that allows researchers to gain an understanding of the tradeoffs and trends for different unit cell

Page 306: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

277

topologies. In addition, while the focus of this work was for L-PBF, ALTO enables consideration

of various AM processes by changing the minimum feature size, build angle restriction, and other

parameters. These contributions and those of others in the community advance our ability to

apply lattice structures to everyday components.

For impact on the broader community, lattice structures form the basis for development

of novel materials and structures with applications across a variety of industries. For example,

consider lattice structures in medical implants. Metal implants are used for their durability and

strength; however, they are often too stiff causing stress shielding and allow limited bone

integration. As referenced in Section 1.1.5, numerous research groups are exploring lattice

structures because they offer the same traditional metallic properties with the opportunity for

reduced stiffness and enhanced bone integration. This research may serve as a stepping stone for

patients to receive custom multi-functional implants, with properties tailored to integrate into

their anatomy and enhance performance. Apart from the medical applications, lattice structures

also offer increased freedom over traditional 3D foams and materials, such as honeycomb

structures for impact absorbing applications. Other applications that benefit the broader

community could include examples such as lightweighting for improving fuel efficiency and

impact absorption for safer cars and planes. Lattice structures offer additional design freedom for

novel material development; systematic design and optimization tools are what is needed to take

advantage of them.

7.4 Recommended Future Research Directions

Based on this work, five recommended research direction are proposed: 1) improving

boundary conditions for homogenization, 2) understanding the impact of as-designed versus as-

printed topology, 3) porosity development in small unit cell sizes (< 2 mm), 4) expansion of

Page 307: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

278

LUCIE, 5) powder removal from AM structures, and 6) continued development of optimization

for additional applications.

As discussed in Chapter 5, the experimental data initially showed significant deviations

from the predicted homogenization properties. The discrepancy is associated with the boundary

conditions for homogenization over constraining and artificially increasing the predicted effective

modulus. Lattice structures have been shown to be sensitive to the number of layers in the testing

configuration. For similar reasons, test coupons are standardized to minimize the edge effects

during testing. Experimental and finite element analysis of large lattice structures to reduce the

boundary effects are expensive to physically and computationally perform, due to the time /cost

of printing and traditional modeling techniques utilized. Further exploration of homogenization

methods, and comparison with other approaches, will benefit DfAM methods such as ALTO so

that they can provide accurate predictions at reduced computational cost.

On the AM side, the practical use of lattice structures has remained limited due to lack of

understanding in performance and build quality. For practical use there are concerns regarding the

reliability of the structures, especially under cyclic or fatigue loading. The majority of

experimental data is for static loading and dynamic loading remains a continuing area of interest

before application of lattice structures can be widespread. Part of the challenge with developing

lattice structure for practical applications is the unknowns of how the as-designed part varies from

the as-printed part. While much is known about surface roughness and anisotropy in non-lattice

components, significantly less information is available for intricate lattice structures. In addition,

these traditional discrepancies between as-designed and as-printed parts are compounded in

lattice structures because the feature size of the discrepancies is often near or similar to the

feature sizes of the lattice. Issues such as surface roughness, sharp corners, anisotropy, dross, and

other geometric variations have a much bigger impact than in non-lattice parts. The focus of this

dissertation regarding unit cell design and optimization was to ensure the manufacturability of the

Page 308: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

279

topology. Future research should address how as-designed parts differ from as-printed parts, the

implications on performance, how to mitigate those discrepancies, microstructures of lattice

struts, qualification strategies, and finishing or post-processing strategies. This practical

understanding of lattice structure performance, combined with systematic optimization

approaches like ALTO, will enable lattice structures in advanced designs.

One example of an issue that arose during CT scanning of powder removability samples

described in Section 5.2.4, was the presence of void regions within the smallest unit cell size. For

the lattice structures with a large unit cell size, these void regions would be expected. After

building the component, the powder would be removed and evacuated from these non-solidified

regions of the lattice through void channels. However, after analysis of the void regions, there

were no channels or pathways from which powder could escape. The CT scan resolution was 15

microns, much smaller than many of the powder particles, and with the exception of a few void

regions near the exterior, there is currently no confirmed explanation of how powder could be

evacuated from these regions. The regularity and spherical nature of the voids is of interest for

understanding of what is causing these void regions as it may relate to other porosity or flaws that

develop in other components. Due to the lack of channels through which powder could be

removed, an assumption is made that these void regions formed during the build process. A few

theories for the void formation are powder particles being drawn away from the region and pulled

into the melt pool on either side or being blown away due to the air flow across the build plate. In

either case, the ability to understand the void formation could lead to improved build reliability in

AM components and lattice structures.

For LUCIE, there is new experimental data available all the time and multiple features

could improve its accessibility and usability. First, as more data becomes available, the database

should be expanded because as additional data will improve the quality of the trends. To support

this effort, a web-based version where researchers can add their own data to the public database,

Page 309: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

280

would improve the sustainability and accessibility of LUCIE. Second, the user interface can

continue to be updated to increase functionality and usability. Third, the ability to directly access

the sources of the data for each data point would provide the user a direct connection between the

data and its source. For example, adding direct access to key process parameters, PBF system,

heat treatment, and finishing processes for each data point or group of data points would helpful

to the user. Fourth, standardization of unit cell names and testing specifications should be

established to make the data reliable and easier to communicate between research groups. Lastly,

in its current state, LUCIE is sufficient to provide mechanical property trends based on the unit

cell topology. Through standardized testing methods, continued additions to the database of key

performance parameters, and incorporation of experimental error or uncertainty into LUCIE, it

will advance beyond simply providing trends to providing exact data based on the specified

volume fraction.

The ability to remove powder from AM components remains is a concern for lattice

structures as well as larger components such as heat sinks. Future research in this area could

include development of predictive tools for powder flow from an AM built component. As a part

of this simulation, thermal simulation to anticipate how cavities or channels will change the

printing process could be used to inform the designer about area way small regions could trap

powder. Generalization of the powder removability factor for larger components and a standard

approach to powder removal would help researchers be able to compare powder removal

techniques and processes to further improve the ability to remove powder. While it remains a

challenge to use full-scale predictive tools in optimization, simplified versions of these full-scale

tools could be paired with ALTO or other topology optimization processes to improve powder

removability.

Finally, the expansion of lattice structures for advanced applications relies on design and

optimization. Regarding expansion of ALTO, there are various research directions that could be

Page 310: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

281

considered such as build orientation consideration, exploration of optimization limits, combined

material and topology optimization, and robust optimization. For this work, the build orientation

is assumed to be vertical according to the z-axis of the ground structure model, however,

changing the build orientation of the lattice would impact the build angle ground structure

restriction, unsupported penalties, and AM-based geometry correction factors. For optimization

limits, additional work could be completed to consider what are the primary limiting factors for

being able to increase the number of design variables that can be optimized in a reasonable time.

For example, the current version of Borg was serial, though a parallelized version of Borg paired

with ALTO could substantially reduce computation time and increase the number of design

variables that can be optimized. Another area where ALTO could be expanded is through

microscale material optimization such as through multi-material unit cell topology optimization,

or simultaneous optimization of the material properties and topology. In this work the microscale

properties are assumed to be fixed and known, however, AM offers opportunities for multiple

materials or changing material properties based on different process parameters which could offer

additional payoffs for unit cell topology optimization. Finally, the field of robust optimization

considers process variability in the optimization problem. Including robust optimization in ALTO

would help the optimal solutions are less susceptible to common flaws or process variation in

AM.

In this work, various optimization objective, constraint, and penalty functions have been

combined to demonstrate a few examples of unit cell generation for differing applications. There

are endless possibilities of other objective functions that could be used to design unit cell

topologies or how the current objectives could be expanded to incorporate other scenarios. For

the powder removability factor, it acts solely on the unit cell topology as opposed to the lattice

structure as a whole. Exploration of how to expand the powder removability factor to account for

the entire lattice would increase the robustness and more accurately represent the actual scenario.

Page 311: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

282

In addition, tortuosity is difficult to solve analytically because the ground structure model has no

consideration for empty space. Therefore, in the ground structure approach there is no model of

the void cavities or channels which gave rise to the method for the powder removability factor.

However, one could consider a correlation between the tortuosity ratio for a lattice structure and

powder removability factor for a unit cell topology in order to understand their correlation.

Another correlation of interest would be to explore if there is any relation between common unit

cell parameters (unit cell size, lattice size, strut diameter, volume fraction, etc.) and the tortuosity

ratio. This could provide insight into powder removal without requiring the powder removability

factor calculation. Another future research direction would be additional exploration of

calculating the tortuosity ratio analytically, possibly through geometric projection methods. As

another example, current objective functions utilize thermal conduction for heat transfer

applications, however, convection plays a critical role in many applications. Another example for

an objective function could be the dynamic response of a lattice structure for impact absorption

scenarios. By expanding the objective functions available, the application of lattice structures

may continue to grow through design and optimization.

Page 312: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

283

References

[1] Howell, L. L., 2001, Compliant Mechanism, Wiley-Interscience.

[2] Kota, S., Lu, K. J., Kreiner, Z., Trease, B., Arenas, J., and Geiger, J., 2005, “Design and

Application of Compliant Mechanisms for Surgical Tools,” J. Biomech. Eng., 127(6), pp.

981–989.

[3] Liu, J., 2012, “Compliant Mechanisms Using Superelastic Nitinol,” Pennsylvania State

University.

[4] Edmondson, B. J., Bowen, L. A., Grames, C. L., Magleby, S. P., Howell, L. L., and

Bateman, T. C., 2016, “Oriceps: Origami-Inspired Forceps,” Proc. ASME SMASIS 2013,

pp. 1–6.

[5] Cronin, J. a., Frecker, M. I., and Mathew, A., 2008, “Design of a Compliant Endoscopic

Suturing Instrument,” J. Med. Device., 2(2), p. 025002.

[6] Aguirre, M. E., 2010, “Design and Optimization of Hybrid Compliant Narrow-Gauge

Surgical Forceps,” Proceedings of the ASME 2010 Conf on SMASIS.

[7] Frecker, M. I., Powell, K. M., and Haluck, R., 2005, “Design of a Multifunctional

Compliant Instrument for Minimally Invasive Surgery.,” J. Biomech. Eng., 127(6), pp.

990–993.

[8] Aguirre, M. E., 2011, “Design and Optimization of Narrow-Gauge Contact-Aided

Compliant Mechanisms for Advanced Minimally Invasive Surgery,” The Pennsylvania

State University.

[9] Lee, C., and Tarbutton, J. A., 2015, “Compliance and Control Characteristics of an

Additive Manufactured-Flexure Stage,” Rev. Sci. Instrum., 86(4).

[10] Cuellar, J. S., Smit, G., Plettenburg, D., and Zadpoor, A., 2018, “Additive Manufacturing

of Non-Assembly Mechanisms,” Addit. Manuf., 21(January), pp. 150–158.

[11] Blanes, C., Mellado, M., and Beltran, P., 2014, “Novel Additive Manufacturing Pneumatic

Actuators and Mechanisms for Food Handling Grippers,” Actuators, 3(3), pp. 205–225.

[12] Mirth, J. A., 2013, “Preliminary Investigations into the Design and Manufacturing of Fully

Compliant Layered Mechanisms,” ASME 2013 International Design Engineering

Technical Conferences & Computers and Information in Engineering Conference.

[13] Gaynor, A. T., Meisel, N. A., Williams, C. B., and Guest, J. K., 2014, “Multiple-Material

Topology Optimization of Compliant Mechanisms Created via Polyjet 3D Printing,” J.

Manuf. Sci. Eng., 136(6), pp. 980–997.

Page 313: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

284

[14] Sakhaei, A. H., Kaijima, S., Lee, T. L., Tan, Y. Y., and Dunn, M. L., 2018, “Design and

Investigation of a Multi-Material Compliant Ratchet-like Mechanism,” Mech. Mach.

Theory, 121, pp. 184–197.

[15] Mirth, J. A., 2016, “The Design and Prototyping of Complex Compliant Mechanisms Via

Multi-Material Additive Manufacturing Techniques,” ASME 2016 International Design

Engineering Technical Conferences and Computers and Information in Engineering

Conference, pp. 1–6.

[16] Merriam, E. G., Jones, J. E., Magleby, S. P., and Howell, L. L., 2013, “Monolithic 2 DOF

Fully Compliant Space Pointing Mechanism,” Mech. Sci., 4(2), pp. 381–390.

[17] Merriam, E. G., Lund, J. M., and Howell, L. L., 2015, “Compound Joints: Behavior and

Benefits of Flexure Arrays,” Precis. Eng., 45, pp. 79–89.

[18] Merriam, E. G., Tolman, K. A., and Howell, L. L., 2016, “Integration of Advanced

Stiffness-Reduction Techniques Demonstrated in a 3D-Printable Joint,” Mech. Mach.

Theory, 105, pp. 260–271.

[19] Merriam, E. G., Jones, J. E., and Howell, L. L., 2014, “Design of 3D-Printed Titanium

Compliant Mechanisms,” Proceedings of the 42nd Aerospace Mechanisms Symposium,

pp. 169–174.

[20] Jovanova, J., and Frecker, M., 2017, “Two Stage Design of Compliant Mechanisms With

Superelastic Compliant,” ASME 2017 Conference on Smar Materials, Adaptive Structures

and Intelligent Systems, pp. 1–9.

[21] Pham, M. T., Teo, T. J., Yeo, S. H., Wang, P., and Nai, M. L. S., 2017, “A 3-D Printed Ti-

6Al-4V 3-DOF Compliant Parallel Mechanism for High Precision Manipulation,”

IEEE/ASME Trans. Mechatronics, 22(5), pp. 2359–2368.

[22] American Cancer Society, 2016, Cancer Facts & Figures 2013, Atlanta, Georgia.

[23] 2015, “Pancreatic Cancer: Treatment Options,” Cancer.Net [Online]. Available:

http://www.cancer.net/cancer-types/pancreatic-cancer/treatment-options.

[24] Greenhalf, W., Grocock, C., Harcus, M., and Neoptolemos, J., 2009, “Screening of High-

Risk Families for Pancreatic Cancer,” Pancreatology, 9(3), pp. 215–222.

[25] Lee, K. S., Sekhar, A., Rofsky, N. M., and Pedrosa, I., 2010, “Prevalence of Incidental

Pancreatic Cysts in the Adult Population on MR Imaging.,” Am. J. Gastroenterol., 105(9),

pp. 2079–84.

[26] Spinelli, K. S., Fromwiller, T. E., Daniel, R. A., Kiely, J. M., Nakeeb, A., Komorowski, R.

A., Wilson, S. D., and Pitt, H. A., 2004, “Cystic Pancreatic Neoplasms: Observe or

Operate.,” Ann. Surg., 239(5), pp. 651–7; discussion 657-9.

[27] Knavel, E. M., and Brace, C. L., 2013, “Tumor Ablation: Common Modalities and

General Practices,” Tech. Vasc. Interv. Radiol., 16(4), pp. 192–200.

Page 314: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

285

[28] Chen, E. L., and Prinz, R. A., 2007, “Long-Term Survival after Pancreatic Cancer

Treatment,” Am. J. Surg., 194(4A), pp. S127–S130.

[29] Werner, J., Combs, S. E., Springfeld, C., Hartwig, W., Hackert, T., and Buchler, M. W.,

2013, “Advanced-Stage Pancreatic Cancer: Therapy Options,” Nat Rev Clin Oncol, 10(6),

pp. 323–333.

[30] Neoptolemos, J. P., Stocken, D. D., Friess, H., Bassi, C., Dunn, J. A., Hickey, H., Beger,

H., Fernandez-Cruz, L., Dervenis, C., Lacaine, F., Falconi, M., Pederzoli, P., Pap, A.,

Spooner, D., Kerr, D. J., and Buchler, M. W., 2004, “A Randomized Trial of

Chemoradiotherapy and Chemotherapy after Resection of Pancreatic Cancer,” N Engl J

Med, 350(12), pp. 1200–1210.

[31] Neoptolemos, J. P., Stocken, D. D., Bassi, C., Ghaneh, P., Cunningham, D., Goldstein, D.,

Padbury, R., Moore, M. J., Gallinger, S., Mariette, C., Wente, M. N., Izbicki, J. R., Friess,

H., Lerch, M. M., Dervenis, C., Olah, A., Butturini, G., Doi, R., Lind, P. A., Smith, D.,

Valle, J. W., Palmer, D. H., Buckels, J. a, Thompson, J., Mckay, C. J., Rawcliffe, C. L.,

and Büchler, M. W., 2010, “Adjuvant Chemotherapy With Fluorouracil Plus Folinic Acid

vs Gemcitabine Following Pancreatic Cancer Resection,” J. Am. Med. Assoc., 304(10),

pp. 1073–1081.

[32] Oettle, H., Post, S., Neuhaus, P., Gellert, K., Langrehr, J., Ridwelski, K., Schramm, H.,

Fahlke, J., Zuelke, C., Burkart, C., Gutberlet, K., Kettner, E., Schmalenberg, H., Weigang-

Koehler, K., Bechstein, W.-O., Niedergethmann, M., Schmidt-Wolf, I., Roll, L., Doerken,

B., and Riess, H., 2007, “Adjuvant Chemotherapy With Gemcitabine vs Observation in

Patients Undergoing Curative-Intent Resection of Pancreatic Cancer,” J. Am. Med.

Assoc., 297(3), pp. 267–277.

[33] Ueno, H., Kosuge, T., Matsuyama, Y., Yamamoto, J., Nakao, A., Egawa, S., Doi, R.,

Monden, M., Hatori, T., Tanaka, M., Shimada, M., and Kanemitsu, K., 2009, “A

Randomised Phase III Trial Comparing Gemcitabine with Surgery-Only in Patients with

Resected Pancreatic Cancer: Japanese Study Group of Adjuvant Therapy for Pancreatic

Cancer.,” Br. J. Cancer, 101(6), pp. 908–15.

[34] Schmidt, J., Abel, U., Debus, J., Harig, S., Hoffmann, K., Herrmann, T., Bartsch, D.,

Klein, J., Mansmann, U., Jäger, D., Capussotti, L., Kunz, R., and Büchler, M. W., 2012,

“Open-Label, Multicenter, Randomized Phase III Trial of Adjuvant Chemoradiation plus

Interferon Alfa-2b versus Fluorouracil and Folinic Acid for Patients with Resected

Pancreatic Adenocarcinoma,” J. Clin. Oncol., 30(33), pp. 4077–4083.

[35] D’Arsonval, A., 1891, “Action Physiologique Des Courants Alternatifs a Grande

Frequence,” Arch Physiol Norm Pathol, 5, pp. 401–408.

[36] Organ, L., 1976, “Electrophysiologic Principles of Radiofrequency Lesion Making,” Appl.

Neurophysiol., 39(2), pp. 69–76.

[37] Krishnamurthy, V. N., Casillas, V. J., and Latorre, L., 2003, “Radiofrequency Ablation of

Hepatic Tumors,” Appl. R, 32(10).

[38] McGahan, J. P., Browning, P. D., Brock, J. M., and Tesluk, H., 1990, “Hepatic Ablation

Page 315: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

286

Using Radiofrequency Electrocautery,” Investig. Radiol., 25(3), pp. 267–270.

[39] Rossi, S., Fornari, F., Pathies, C., and Buscarini, L., 1990, “Thermal Lesions Induced by

480 KHz Localized Current Field in Guinea Pig and Pig Liver,” Tumori, 76(1), pp. 54–57.

[40] VanSonnenberg, E., McMullen, W., and Solbiati, L., 2010, Tumor Ablation, Springer New

York.

[41] Hong, K., and Georgiades, C., 2010, “Radiofrequency Ablation: Mechanism of Action

and Devices,” J. Vasc. Interv. Radiol., 21(SUPPL. 8), pp. S179–S186.

[42] Dewhirst, M. W., Viglianti, B. L., Lora-Michiels, M., Hanson, M., and Hoopes, P. J.,

2003, “Basic Principles of Thermal Dosimetry and Thermal Thresholds for Tissue

Damage from Hyperthermia,” Int. J. Hyperth., 19(3), pp. 267–294.

[43] Sapareto, S. A., and Dewey, W. C., 1984, “Thermal Dose Determination in Cancer

Therapy,” Int. J. Radiat. Oncol. Biol. Phys., 10(6), pp. 787–800.

[44] Ryan, T. P., Kwok, J., and Beetel, R., 2003, “Simulations of Percutaneous RF Ablation

Systems,” SPIE Conference Proceedings, pp. 71–88.

[45] Brace, C. L., 2009, “Radiofrequency and Microwave Ablation of the Liver, Lung, Kidney

and Bone: What Are the Differences,” Curr. Probl. Diagn. Radiol., 38(3), pp. 135–143.

[46] Feldman, L. S., Fuchshuber, P. R., and Editors, D. B. J., 2012, The SAGES Manual on the

Fundamental Use of Surgical Energy (FUSE), Springer-Verlag, New York.

[47] Dewhirst, M. W., Vujaskovic, Z., Jones, E., and Thrall, D., 2005, “Re-Setting the Biologic

Rationale for Thermal Therapy.,” Int. J. Hyperthermia, 21(8), pp. 779–90.

[48] Nemcek, A. a., 2006, “Complications of Radiofrequency Ablation of Neoplasms,” Semin.

Intervent. Radiol., 23(2), pp. 177–187.

[49] Lu, D. S. K., Raman, S. S., Limanond, P., Aziz, D., Economou, J., Busuttil, R., and Sayre,

J., 2003, “Influence of Large Peritumoral Vessels on Outcome of Radiofrequency

Ablation of Liver Tumors,” J. Vasc. Interv. Radiol., 14(10), pp. 1267–1274.

[50] Saldanha, D. F., Khiatani, V. L., Carrillo, T. C., Yap, F. Y., Bui, J. T., Knuttinen, M. G.,

Owens, C. A., and Gaba, R. C., 2010, “Current Tumor Ablation Technologies: Basic

Science and Device Review,” Semin. Intervent. Radiol., 27(3), pp. 247–254.

[51] Mulier, S., Miao, Y., Mulier, P., Dupas, B., Pereira, P., De Baere, T., Lencioni, R.,

Leveillee, R., Marchal, G., Michel, L., and Ni, Y., 2006, “Electrodes and Multiple

Electrode Systems for Radio Frequency Ablation: A Proposal for Updated Terminology,”

Eur. J. Radiol., 15, pp. 57–73.

[52] D’Onofrio, M., Barbi, E., Girelli, R., Martone, E., Gallotti, A., Salvia, R., Martini, P.-T.,

Bassi, C., Pederzoli, P., and Pozzi Mucelli, R., 2010, “Radiofrequency Ablation of Locally

Advanced Pancreatic Adenocarcinoma: An Overview.,” World J. Gastroenterol., 16(28),

pp. 3478–3483.

Page 316: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

287

[53] Varghese, T., Techavipoo, U., Zagzebski, J. a, and Lee, F. T., 2004, “Impact of Gas

Bubbles Generated during Interstitial Ablation on Elastographic Depiction of in Vitro

Thermal Lesions.,” J. Ultrasound Med., 23(4), pp. 535–544; quiz 545–546.

[54] Raman, S. S., Lu, D. S. K. K., Vodopich, D. J., Sayre, J., and Lassman, C., 2000,

“Creation of Radiofrequency Lesions in a Porcine Model: Correlation with Sonography,

CT, and Histopathology,” Am. J. Roentgenol., 175(5), pp. 1253–1258.

[55] Leyendecker, J. R., Dodd, Gerald D, I., Halff, G. A., Mccoy, V. A., Napier, D. H.,

Hubbard, L. G., Chintapalli, K. N., Chopra, S., Washburn, W. K., Esterl, R. M., Cigarroa,

F. G., Kohlmeier, R. E., and Sharkey, F. E., 2002, “Sonographically Observed Echogenic

Response During Intraoperative Radiofrequency Ablation of Cirrhotic Livers : Pathologic

Correlation,” Am. J. Roentgenol., 178(May), pp. 1147–1151.

[56] Cha, C. H., Lee, F. T., Gurne, J. M., Markhardt, B. K., Warner, T. F., Kelcz, F., and

Mahvi, D. M., 2000, “CT Versus Monitoring Ablation Sonography for Radiofrequency in

a Porcine Liver,” Imaging, 175(3), pp. 705–711.

[57] Wu, H., Wilkins, L. R., Ziats, N. P., Haaga, J. R., and Exner, A. A., 2014, “Real-Time

Monitoring of Radiofrequency Ablation and Postablation Assessment: Accuracy of

Contrast-Enhanced US in Experimental Rat Liver Model.,” Radiology, 270(1), pp. 107–

16.

[58] Isobe, Y., Watanabe, H., Yamazaki, N., Lu, X., Kobayashi, Y., Miyashita, T., Hashizume,

M., and Fujie, M. G., 2013, “Real-Time Temperature Control System Based on the Finite

Element Method for Liver Radiofrequency Ablation: Effect of the Time Interval on

Control.,” International Conference of the IEEE Engineering in Medicine and Biology

Society, pp. 392–6.

[59] Jin, C., He, Z., and Liu, J., 2014, “MRI-Based Finite Element Simulation on

Radiofrequency Ablation of Thyroid Cancer,” Comput. Methods Programs Biomed.,

113(2), pp. 529–538.

[60] Zhang, M., Zhou, Z., Wu, S., Lin, L., Gao, H., and Feng, Y., 2015, “Simulation of

Temperature Field for Temperature-Controlled Radio Frequency Ablation Using a

Hyperbolic Bioheat Equation and Temperature-Varied Voltage Calibration: A Liver-

Mimicking Phantom Study,” Phys. Med. Biol., 60(24), pp. 9455–9471.

[61] Chang, I., 2003, “Finite Element Analysis of Hepatic Radiofrequency Ablation Probes

Using Temperature-Dependent Electrical Conductivity.,” Biomed. Eng. Online, 2, p. 12.

[62] Schutt, D. J., and Haemmerich, D., 2008, “Effects of Variation in Perfusion Rates and of

Perfusion Models in Computational Models of Radio Frequency Tumor Ablation.,” Med.

Phys., 35(8), pp. 3462–3470.

[63] Solazzo, S. a, Liu, Z., Lobo, S. M., Ahmed, M., Hines-Peralta, A. U., Lenkinski, R. E.,

and Goldberg, S. N., 2005, “Radiofrequency Ablation: Importance of Background Tissue

Electrical Conductivity--an Agar Phantom and Computer Modeling Study.,” Radiology,

236(2), pp. 495–502.

Page 317: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

288

[64] Duan, B., Wen, R., Fu, Y., Chua, K. J., and Chui, C. K., 2016, “Probabilistic Finite

Element Method for Large Tumor Radiofrequency Ablation Simulation and Planning,”

Med. Eng. Phys., 38(11), pp. 1360–1368.

[65] Subramanian, S., and Mast, T. D., 2015, “Optimization of Tissue Physical Parameters for

Accurate Temperature Estimation from Finite-Element Simulation of Radiofrequency

Ablation,” Phys. Med. Biol., 60(19), pp. N345–N355.

[66] Huang, W., and Chui, C., 2012, “A Radio-Frequency Ablation Planning System Using

Stochastic Finite Element Method,” IEEE/SCIE.

[67] Tungjitkusolmun, S., Staelin, S. T., Haemmerich, D., Tsai, J. Z., Cao, H., Webster, J. G.,

Lee, F. T., Mahvi, D. M., and Vorperian, V. R., 2002, “Three-Dimensional Finite-Element

Analyses for Radio-Frequency Hepatic Tumor Ablation,” IEEE Trans. Biomed. Eng.,

49(1), pp. 3–9.

[68] Lim, D., Namgung, B., Woo, D. G., Choi, J. S., Kim, H. S., and Tack, G. R., 2010, “Effect

of Input Waveform Pattern and Large Blood Vessel Existence on Destruction of Liver

Tumor Using Radiofrequency Ablation: Finite Element Analysis.,” J. Biomech. Eng.,

132(6), p. 061003.

[69] Haemmerich, D., Wright, A. W., Mahvi, D. M., Lee, F. T., and Webster, J. G., 2003,

“Hepatic Bipolar Radiofrequency Ablation Creates Coagulation Zones Close to Blood

Vessels: A Finite Element Study,” Med. Biol. Eng. Comput., 41(3), pp. 317–323.

[70] Dos Santos, I., Haemmerich, D., Schutt, D., Da Rocha, A. F., and Menezes, L. R., 2009,

“Probabilistic Finite Element Analysis of Radiofrequency Liver Ablation Using the

Unscented Transform.,” Phys. Med. Biol., 54(3), pp. 627–640.

[71] Chang, I. a, and Nguyen, U. D., 2004, “Thermal Modeling of Lesion Growth with

Radiofrequency Ablation Devices.,” Biomed. Eng. Online, 3(27).

[72] Barauskas, R., Gulbinas, A., Vanagas, T., and Barauskas, G., 2008, “Finite Element

Modeling of Cooled-Tip Probe Radiofrequency Ablation Processes in Liver Tissue,”

Comput. Biol. Med., 38(6), pp. 694–708.

[73] Altrogge, I., Preusser, T., Kröger, T., Büskens, C., Pereira, P. L., Schmidt, D., and

Peitgen, H. O., 2007, “Multiscale Optimization of the Probe Placement for

Radiofrequency Ablation,” Acad. Radiol., 14(11), pp. 1310–1324.

[74] Haemmerich, D., Tungjitkusolmun, S., Staelin, S. T., Lee, F. T., Mahvi, D. M., and

Webster, J. G., 2002, “Finite-Element Analysis of Hepatic Multiple Probe Radio-

Frequency Ablation,” IEEE Trans. Biomed. Eng., 49(8), pp. 836–842.

[75] Haemmerich, D., and Webster, J. G., 2005, “Automatic Control of Finite Element Models

for Temperature-Controlled Radiofrequency Ablation,” Biomed. Eng. Online, 4, pp. 1–8.

[76] Varadarajulu, S., Jhala, N. C., and Drelichman, E. R., 2009, “EUS-Guided

Radiofrequency Ablation with a Prototype Electrode Array System in an Animal Model

(with Video),” Gastrointest. Endosc., 70(2), pp. 372–376.

Page 318: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

289

[77] Gaidhane, M., Smith, I., Ellen, K., Gatesman, J., Habib, N., Foley, P., Moskaluk, C., and

Kahaleh, M., 2012, “Endoscopic Ultrasound-Guided Radiofrequency Ablation (EUS-

RFA) of the Pancreas in a Porcine Model,” Gastroenterol. Res. Pract., 2012, pp. 3–8.

[78] Pai, M., Habib, N., Senturk, H., Lakhtakia, S., Reddy, N., Cicinnati, V. R., Kaba, I.,

Beckebaum, S., Drymousis, P., Kahaleh, M., and Brugge, W., 2015, “Endoscopic

Ultrasound Guided Radiofrequency Ablation, for Pancreatic Cystic Neoplasms and

Neuroendocrine Tumors,” World J. Gastrointest. Surg., 7(4), pp. 52–59.

[79] Bas-Cutrina, F., Bargalló, D., and Gornals, J. B., 2017, “Small Pancreatic Insulinoma:

Successful Endoscopic Ultrasound-Guided Radiofrequency Ablation in a Single Session

Using a 22-G Fine Needle,” Dig. Endosc., 29(5), pp. 636–638.

[80] Chaudhary, S., and Sun, S. Y., 2017, “Endoscopic Ultrasound-Guided Radiofrequency

Ablation in Gastroenterology: New Horizons in Search,” World J. Gastroenterol., 23(27),

pp. 4892–4896.

[81] Choi, J. H., Seo, D. W., Song, T. J., Park, D. H., Lee, S. S., Lee, S. K., and Kim, M. H.,

2018, “Endoscopic Ultrasound-Guided Radiofrequency Ablation for Management of

Benign Solid Pancreatic Tumors,” Endoscopy, 50(11), pp. 1099–1104.

[82] Barthet, M., Giovannini, M., Lesavre, N., Boustiere, C., Napoleon, B., Koch, S., Gasmi,

M., Vanbiervliet, G., and Gonzalez, J. M., 2019, “Endoscopic Ultrasound-Guided

Radiofrequency Ablation for Pancreatic Neuroendocrine Tumors and Pancreatic Cystic

Neoplasms: A Prospective Multicenter Study,” Endoscopy, 51(9), pp. 836–842.

[83] Oleinikov, K., Dancour, A., Epshtein, J., Benson, A., Mazeh, H., Tal, I., Matalon, S.,

Benbassat, C. A., Livovsky, D. M., Goldin, E., Gross, D. J., Jacob, H., and Grozinsky-

Glasberg, S., 2019, “Endoscopic Ultrasound-Guided Radiofrequency Ablation: A New

Therapeutic Approach for Pancreatic Neuroendocrine Tumors,” J. Clin. Endocrinol.

Metab., 104(7), pp. 2637–2647.

[84] Ofosu, A., Ramai, D., and Adler, D. G., 2019, “Endoscopic Ultrasound-Guided Ablation

of Pancreatic Cystic Neoplasms: Ready for Prime Time?,” Ann. Gastroenterol., 32(1), pp.

39–45.

[85] Waung, J. A., Todd, J. F., Keane, M. G., and Pereira, S. P., 2016, “Successful

Management of a Sporadic Pancreatic Insulinoma by Endoscopic Ultrasound-Guided

Radio-Frequency Ablation,” Endoscopy, 48, pp. E144–E145.

[86] Carrara, S., Arcidiacono, P. G., Albarello, L., Addis, A., Enderle, M. D., Boemo, C.,

Campagnol, M., Ambrosi, A., Doglioni, C., and Testoni, P. A., 2008, “Endoscopic

Ultrasound-Guided Application of a New Hybrid Cryotherm Probe in Porcine Pancreas: A

Preliminary Study,” Endoscopy, 40, pp. 321–326.

[87] Arcidiacono, P. G., Carrara, S., Reni, M., Petrone, M. C., Cappio, S., Balzano, G., Boemo,

C., Cereda, S., Nicoletti, R., Enderle, M. D., Neugebauer, A., Von Renteln, D., Eickhoff,

A., and Testoni, P. A., 2012, “Feasibility and Safety of EUS-Guided Cryothermal

Ablation in Patients with Locally Advanced Pancreatic Cancer,” Gastrointest. Endosc.,

76(6), pp. 1142–1151.

Page 319: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

290

[88] Gao, W., Zhang, Y., Ramanujan, D., Ramani, K., Chen, Y., Williams, C. B., Wang, C. C.

L., Shin, Y. C., Zhang, S., and Zavattieri, P. D., 2015, “The Status, Challenges, and Future

of Additive Manufacturing in Engineering,” Comput. Des., 69, pp. 65–89.

[89] Wong, K. V., and Hernandez, A., 2012, “A Review of Additive Manufacturing,” ISRN

Mech. Eng., 2012, pp. 1–10.

[90] Guo, N., and Leu, M. C., 2013, “Additive Manufacturing: Technology, Applications and

Research Needs,” Front. Mech. Eng., 8(3), pp. 215–243.

[91] Abdulhameed, O., Al-Ahmari, A., Ameen, W., and Mian, S. H., 2019, “Additive

Manufacturing: Challenges, Trends, and Applications,” Adv. Mech. Eng., 11(2), pp. 1–27.

[92] Cali, J., Calian, D. A., Amati, C., Kleinberger, R., Steed, A., Kautz, J., and Weyrich, T.,

2012, “3D-Printing of Non-Assembly, Articulated Models,” ACM Trans. Graph., 31(6).

[93] Mohammed, M. I., P. Fitzpatrick, A., and Gibson, I., 2017, “Customised Design of a

Patient Specific 3D Printed Whole Mandible Implant,” KnE Eng., 2(2), p. 104.

[94] Tuck, C. J., Hague, R. J. M., Ruffo, M., Ransley, M., and Adams, P., 2008, “Rapid

Manufacturing Facilitated Customization,” Int. J. Comput. Integr. Manuf., 21(3), pp. 245–

258.

[95] Maji, P. K., Banerjee, A. J., Banerjee, P. S., and Karmakar, S., 2014, “Additive

Manufacturing in Prosthesis Development – a Case Study,” Rapid Prototyp. J., 20(6), pp.

480–489.

[96] Schmelzle, J., Kline, E. V., Dickman, C. J., Reutzel, E. W., Jones, G., and Simpson, T.

W., 2015, “(Re)Designing for Part Consolidation: Understanding the Challenges of Metal

Additive Manufacturing,” J. Mech. Des., 137(11), p. 111404.

[97] Kellner, T., 2017, “An Epiphany Of Disruption: GE Additive Chief Explains How 3D

Printing Will Upend Manufacturing,” GE Reports [Online]. Available:

https://www.ge.com/reports/epiphany-disruption-ge-additive-chief-explains-3d-printing-

will-upend-manufacturing/. [Accessed: 24-May-2018].

[98] MacDonald, E., and Wicker, R., 2016, “Multiprocess 3D Printing for Increasing

Component Functionality,” Science (80-. )., 353(6307).

[99] Kataria, A., and Rosen, D. W., 2001, “Building around Inserts: Methods for Fabricating

Complex Devices in Stereolithography,” Rapid Prototyp. J., 7(5), pp. 253–262.

[100] Meisel, N. A., Elliott, A. M., and Williams, C. B., 2015, “A Procedure for Creating

Actuated Joints via Embedding Shape Memory Alloys in PolyJet 3D Printing,” J. Intell.

Mater. Syst. Struct., 26(12), pp. 1498–1512.

[101] Ota, H., Emaminejad, S., Gao, Y., Zhao, A., Wu, E., Challa, S., Chen, K., Fahad, H. M.,

Jha, A. K., Kiriya, D., Gao, W., Shiraki, H., Morioka, K., Ferguson, A. R., Healy, K. E.,

Davis, R. W., and Javey, A., 2016, “Application of 3D Printing for Smart Objects with

Embedded Electronic Sensors and Systems,” Adv. Mater. Technol., 1(1).

Page 320: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

291

[102] Doubrovski, E. L., Tsai, E. Y., Dikovsky, D., Geraedts, J. M. P., Herr, H., and Oxman, N.,

2015, “Voxel-Based Fabrication through Material Property Mapping: A Design Method

for Bitmap Printing,” CAD Comput. Aided Des., 60, pp. 3–13.

[103] Garland, A., and Fadel, G., 2015, “Design and Manufacturing Functionally Gradient

Material Objects With an Off the Shelf Three-Dimensional Printer: Challenges and

Solutions,” J. Mech. Des., 137(11), p. 111407.

[104] Everton, S. K., Hirsch, M., Stravroulakis, P., Leach, R., and Clare, A. T., 2016, “Review

of in Situ Process Monitoring and in Situ Metrology for Metal Additive Manufacturing,”

Mater. Des., 95, pp. 431–445.

[105] Chua, Z. Y., Ahn, I. H., and Moon, S. K., 2017, “Process Monitoring and Inspection

Systems in Metal Additive Manufacturing: Status and Applications,” Int. J. Precis. Eng.

Manuf. - Green Technol., 4(2), pp. 235–245.

[106] Grasso, M., and Colosimo, B. M., 2017, “Process Defects and in Situ Monitoring Methods

in Metal Powder Bed Fusion: A Review,” Meas. Sci. Technol., 28(4).

[107] Montazeri, M., and Rao, P., 2017, “In-Process Condition Monitoring in Laser Powder Bed

Fusion (LPBF),” pp. 1264–1278.

[108] Repossini, G., Laguzza, V., Grasso, M., and Colosimo, B. M., 2017, “On the Use of

Spatter Signature for In-Situ Monitoring of Laser Powder Bed Fusion,” Addit. Manuf., 16,

pp. 35–48.

[109] Zhang, B., Land, W. S., Ziegert, J., Davies, A., Ii, W. S. L., Ziegert, J., and Davies, A.,

2015, “In Situ Monitoring of Laser Powder Bed Fusion Additive Manufacturing Using

Digital Fringe Projection Technique,” Proc. ASPE 2015 Spring Top. Meet., (January), pp.

47–52.

[110] Zhao, C., Fezzaa, K., Cunningham, R. W., Wen, H., De Carlo, F., Chen, L., Rollett, A. D.,

and Sun, T., 2017, “Real-Time Monitoring of Laser Powder Bed Fusion Process Using

High-Speed X-Ray Imaging and Diffraction,” Sci. Rep., 7(1), pp. 1–11.

[111] Yadroitsev, I., Krakhmalev, P., and Yadroitsava, I., 2014, “Selective Laser Melting of

Ti6Al4V Alloy for Biomedical Applications: Temperature Monitoring and

Microstructural Evolution,” J. Alloys Compd., 583, pp. 404–409.

[112] Frazier, W. E., 2014, “Metal Additive Manufacturing: A Review,” J. Mater. Eng.

Perform., 23(6), pp. 1917–1928.

[113] Bhavar, V., Kattire, P., Patil, V., Khot, S., Gujar, K., and Singh, R., 2014, “A Review on

Powder Bed Fusion Technology of Metal Additive Manufacturing,” Mater. Sci. Technol.

2014.

[114] Herzog, D., Seyda, V., Wycisk, E., and Emmelmann, C., 2016, “Additive Manufacturing

of Metals,” Acta Mater., 117, pp. 371–392.

[115] Fielding, J. D., Gorham, R., Davis, A., Bouffard, B., and Morris, K., 2016, Department of

Page 321: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

292

Defense Joint Additive Manufacturing Roadmap.

[116] Dong, G., Tang, Y., and Zhao, Y. F., 2017, “A Survey of Modeling of Lattice Structures

Fabricated by Additive Manufacturing,” J. Mech. Des., 139(10), pp. 1–13.

[117] Chu, C., Graf, G., and Rosen, D. W., 2008, “Design for Additive Manufacturing of

Cellular Structures,” Comput. Aided. Des. Appl., 5(5), pp. 686–696.

[118] Seepersad, C. C., Allen, J. K., McDowell, D. L., and Mistree, F., 2008, “Multifunctional

Topology Design of Cellular Material Structures,” J. Mech. Des., 130.

[119] Gibson, I., Rosen, D. W., and Stucker, B., 2015, “Directed Energy Deposition,” Additive

Manufacturing Technologies: 3D Printing, Rapid Prototyping, and Direct Digital

Manufacturing, pp. 245–268.

[120] Thompson, M. K., Moroni, G., Vaneker, T., Fadel, G., Campbell, R. I., Gibson, I.,

Bernard, A., Schulz, J., Graf, P., Ahuja, B., and Martina, F., 2016, “Design for Additive

Manufacturing: Trends, Opportunities, Considerations, and Constraints,” CIRP Ann. -

Manuf. Technol., 65(2), pp. 737–760.

[121] ASTM International, 2017, ISO/ASTM 52910:2017 Standard Guidelines for Design for

Additive Manufacturing.

[122] Wiberg, A., Persson, J., and Ölvander, J., 2019, “Design for Additive Manufacturing – a

Review of Available Design Methods and Software,” Rapid Prototyp. J., 25(6), pp. 1080–

1094.

[123] Laverne, F., Segonds, F., Anwer, N., and Le Coq, M., 2015, “Assembly Based Methods to

Support Product Innovation in Design for Additive Manufacturing: An Exploratory Case

Study,” J. Mech. Des., 137(12), p. 121701.

[124] Yang, S., and Zhao, Y. F., 2015, “Additive Manufacturing-Enabled Design Theory and

Methodology: A Critical Review,” Int. J. Adv. Manuf. Technol., 80(1–4), pp. 327–342.

[125] Altair, “Optistruct” [Online]. Available: https://www.altair.com/optistruct/.

[126] Autodesk, 2019, “Netfabb Ultimate 2019” [Online]. Available:

https://www.autodesk.com/products/netfabb/overview.

[127] Doubrovski, Z., Verlinden, J. C., and Geraedts, J. M. P., 2011, “OPTIMAL DESIGN FOR

ADDITIVE MANUFACTURING: OPPORTUNITIES AND CHALLENGES,” ASME

2011 International Design Engineering Technical Conferences & Computers and

Information in Engineering Conference, pp. 1–12.

[128] Thomas, D., 2009, “The Development of Design Rules for Selective Laser Melting The

Development of Design Rules for Selective Laser Melting.”

[129] Vayre, B., Vignat, F., and Villeneuve, F., 2012, “Designing for Additive Manufacturing,”

Procedia CIRP, 3(1), pp. 632–637.

Page 322: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

293

[130] ASTM International, 2019, Additive Manufacturing — Design — Part 1 : Laser-Based

Powder Bed Fusion of Metals 1 (ISO/ASTM 52911-1).

[131] ASTM International, 2019, Additive Manufacturing — Technical Design Guideline for

Powder Bed Fusion Part 2: Laser-Based Powder Bed Fusion of Polymers (ISO/ASTM

52911-2).

[132] ASTM International, 2019, Additive Manufacturing - Design - Requirements, Guidelines

and Recommendations (ISO/ASTM52910).

[133] Chen, W., Watts, S., Jackson, J. A., Smith, W. L., Tortorelli, D. A., and Spadaccini, C. M.,

2019, “Stiff Isotropic Lattices beyond the Maxwell Criterion,” Sci. Adv., 5(9), pp. 1–7.

[134] Pham, M.-S., Liu, C., Todd, I., and Lertthanasarn, J., 2019, “Damage-Tolerant Architected

Materials Inspired by Crystal Microstructure,” Nature.

[135] Tang, Y., and Zhao, Y. F., 2016, “A Survey of the Design Methods for Additive

Manufacturing to Improve Functional Performance,” Rapid Prototyp. J., 22(3), pp. 569–

590.

[136] Gibson, L. J., and Ashby, M. F., 1997, Cellular Solids, Cambridge University Press,

Cambridge, New York.

[137] Ashby, M. F., Evans, A. G., Fleck, N. A., Gibson, L. J., Hutchinson, J. W., and Wadley,

H., 2000, Metal Foams: A Design Guide, Butterworth-Heinemann, Boston, MA.

[138] Ashby, M. ., 2006, “The Properties of Foams and Lattices,” Philos. Trans. R. Soc. A

Math. Phys. Eng. Sci., 364(1838), pp. 15–30.

[139] Gibson, L. J., and Ashby, M. F., 1982, “The Mechanics of Three-Dimensional Cellular

Materials,” Proc. R. Soc. London, 382(1782), pp. 305–331.

[140] Wadley, H., Fleck, N. A., and Evans, A. G., 2003, “Fabrication and Structural

Performance of Periodic Cellular Metal Sandwich Structures,” Compos. Sci. Technol.,

63(16), pp. 2331–2343.

[141] Wadley, H. N. ., 2006, “Multifunctional Periodic Cellular Metals,” Philos. Trans. R. Soc.

A Math. Phys. Eng. Sci., 364(1838), pp. 31–68.

[142] Queheillalt, D. T., and Wadley, H., 2011, “Hollow Pyramidal Lattice Truss Structures,”

Int. J. Mater. Res., 102(4).

[143] Queheillalt, D. T., Murty, Y., and Wadley, H., 2008, “Mechanical Properties of an

Extruded Pyramidal Lattice Truss Sandwich Structure,” Scr. Mater., 58(1), pp. 76–79.

[144] Queheillalt, D. T., and Wadley, H., 2005, “Pyramidal Lattice Truss Structures with

Hollow Trusses,” Mater. Sci. Eng. A, 397(1–2), pp. 132–137.

[145] Queheillalt, D. T., and Wadley, H., 2005, “Cellular Metal Lattices with Hollow Trusses,”

Acta Mater., 53(2), pp. 303–313.

Page 323: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

294

[146] Queheillalt, D. T., and Wadley, H., 2009, “Titanium Alloy Lattice Truss Structures,”

Mater. Des., 30(6), pp. 1966–1975.

[147] Wadley, H., Masta, M. R. O., Dharmasena, K. P., Compton, B. G., Gamble, E. A., and

Zok, F. W., 2013, “Effect of Core Topology on Projectile Penetration in Hybrid

Aluminum / Alumina Sandwich Structures Q,” Int. J. Impact Eng., 62, pp. 99–113.

[148] Wadley, H., Dharmasena, K. P., O’Masta, M. R., and Wetzel, J. J., 2013, “Impact

Response of Aluminum Corrugated Core Sandwich Panels,” Int. J. Impact Eng., 62, pp.

114–128.

[149] Yungwirth, C. J., Radford, D. D., Aronson, M., and Wadley, H., 2008, “Experiment

Assessment of the Ballistic Response of Composite Pyramidal Lattice Truss Structures,”

Compos. Part B Eng., 39(3), pp. 556–569.

[150] Dong, L., Deshpande, V., and Wadley, H., 2015, “Mechanical Response of Ti–6Al–4V

Octet-Truss Lattice Structures,” Int. J. Solids Struct., 60–61, pp. 107–124.

[151] Dong, L., King, W. P., Raleigh, M., and Wadley, H., 2018, “A Microfabrication Approach

for Making Metallic Mechanical Metamaterials,” Mater. Des., 160, pp. 147–168.

[152] Holloman, R. L., Deshpande, V., Hanssen, A. G., Fleming, K. M., Scully, J. R., and

Wadley, H., 2013, “Tubular Aluminum Cellular Structures: Fabrication and Mechanical

Response,” J. Mech. Mater. Struct., 8(1), pp. 65–94.

[153] Kooistra, G. W., Deshpande, V. S., and Wadley, H., 2004, “Compressive Behavior of Age

Hardenable Tetrahedral Lattice Truss Structures Made from Aluminium,” Acta Mater.,

52(14), pp. 4229–4237.

[154] Kooistra, G. W., and Wadley, H., 2007, “Lattice Truss Structures from Expanded Metal

Sheet,” Mater. Des., 28(2), pp. 507–514.

[155] Moongkhamklang, P., Deshpande, V. S., and Wadley, H., 2010, “The Compressive and

Shear Response of Titanium Matrix Composite Lattice Structures,” Acta Mater., 58(8),

pp. 2822–2835.

[156] Moongkhamklang, P., and Wadley, H., 2010, “Titanium Alloy Lattice Structures with

Millimeter Scale Cell Sizes,” Adv. Eng. Mater., 12(11), pp. 1111–1116.

[157] O’Masta, M. R., Dong, L., St-Pierre, L., Wadley, H., and Deshpande, V. S., 2017, “The

Fracture Toughness of Octet-Truss Lattices,” J. Mech. Phys. Solids, 98, pp. 271–289.

[158] Rosen, D. W., Johnston, S., and Reed, M., 2006, “Design of General Lattice Structures for

Lightweight and Compliance Applications,” Rapid Manuf. Conf.

[159] Rosen, D. W., 2007, “Computer-Aided Design for Additive Manufacturing of Cellular

Structures,” Comput. Aided. Des. Appl., 4(1–6), pp. 585–594.

[160] Helou, M., and Kara, S., 2017, “Design, Analysis and Manufacturing of Lattice

Structures: An Overview,” Int. J. Comput. Integr. Manuf., 31(3), pp. 243–261.

Page 324: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

295

[161] Nguyen, J., Park, S., and Rosen, D. W., 2013, “Heuristic Optimization Method for

Cellular Structure Design of Light Weight Components,” Int. J. Precis. Eng. Manuf.,

14(6), pp. 1071–1078.

[162] Cheng, L., Bai, J., and To, A. C., 2019, “Functionally Graded Lattice Structure Topology

Optimization for the Design of Additive Manufactured Components with Stress

Constraints,” Comput. Methods Appl. Mech. Eng., 344, pp. 334–359.

[163] Boccaccio, A., Uva, A. E., Fiorentino, M., Mori, G., and Monno, G., 2016, “Geometry

Design Optimization of Functionally Graded Scaffolds for Bone Tissue Engineering: A

Mechanobiological Approach,” PLoS One, 11(1).

[164] An, J., Teoh, J. E. M., Suntornnond, R., and Chua, C. K., 2015, “Design and 3D Printing

of Scaffolds and Tissues,” Engineering, 1(2), pp. 261–268.

[165] Yeong, W.-Y., Chua, C.-K., Leong, K.-F., and Chandrasekaran, M., 2004, “Rapid

Prototyping in Tissue Engineering: Challenges and Potential,” Trends Biotechnol., 22(12),

pp. 643–652.

[166] Warnke, P. H., Douglas, T., Wollny, P., Sherry, E., Steiner, M., Galonska, S., Becker, S.

T., Springer, I. N., Wiltfang, J., and Sivananthan, S., 2009, “Rapid Prototyping: Porous

Titanium Alloy Scaffolds Produced by Selective Laser Melting for Bone Tissue

Engineering,” Tissue Eng. Part C Methods, 15(2), pp. 115–124.

[167] Leong, K. F., Chua, C. K., Sudarmadji, N., and Yeong, W. Y., 2008, “Engineering

Functionally Graded Tissue Engineering Scaffolds,” J. Mech. Behav. Biomed. Mater.,

1(2), pp. 140–152.

[168] Hollister, S. J., 2005, “Porous Scaffold Design for Tissue Engineering,” Nat. Mater., 4(7),

pp. 518–524.

[169] Dias, M. R., Guedes, J. M., Flanagan, C. L., Hollister, S. J., and Fernandes, P. R., 2014,

“Optimization of Scaffold Design for Bone Tissue Engineering: A Computational and

Experimental Study,” Med. Eng. Phys., 36(4), pp. 448–457.

[170] Chua, C. K., Leong, K. F., Cheah, C. M., and Chua, S. W., 2003, “Development of a

Tissue Engineering Scaffold Structure Library for Rapid Prototyping. Part 1: Investigation

and Classification,” Int. J. Adv. Manuf. Technol., 21(4), pp. 291–301.

[171] Taniguchi, N., Fujibayashi, S., Takemoto, M., Sasaki, K., Otsuki, B., Nakamura, T.,

Matsushita, T., Kokubo, T., and Matsuda, S., 2016, “Effect of Pore Size on Bone Ingrowth

into Porous Titanium Implants Fabricated by Additive Manufacturing: An in Vivo

Experiment,” Mater. Sci. Eng. C, 59, pp. 690–701.

[172] Hanks, B., Dinda, S., and Joshi, S., 2018, “Redesign of the Femoral Stem for a Total Hip

Arthroplasty for Additive Manufacturing,” ASME International Design Engineering

Technical Conferences and Computers and Information in Engineering Conference, pp.

1–13.

[173] Burton, H. E., Eisenstein, N. M., Lawless, B. M., Jamshidi, P., Segarra, M. A., Addison,

Page 325: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

296

O., Shepherd, D. E. T., Attallah, M. M., Grover, L. M., and Cox, S. C., 2019, “The Design

of Additively Manufactured Lattices to Increase the Functionality of Medical Implants,”

Mater. Sci. Eng. C, 94(September 2018), pp. 901–908.

[174] Jonitz-Heincke, A., Wieding, J., Schulze, C., Hansmann, D., and Bader, R., 2013,

“Comparative Analysis of the Oxygen Supply and Viability of Human Osteoblasts in

Three-Dimensional Titanium Scaffolds Produced by Laser-Beam or Electron-Beam

Melting,” Materials (Basel)., 6(11), pp. 5398–5409.

[175] Chua, C. K., Leong, K. F., Cheah, C. M., and Chua, S. W., 2003, “Development of a

Tissue Engineering Scaffold Structure Library for Rapid Prototyping. Part 2: Parametric

Library and Assembly Program,” Int. J. Adv. Manuf. Technol., 21(4), pp. 302–312.

[176] Tan, X. P., Tan, Y. J., Chow, C. S. L., Tor, S. B., and Yeong, W. Y., 2017, “Metallic

Powder-Bed Based 3D Printing of Cellular Scaffolds for Orthopaedic Implants: A State-

of-the-Art Review on Manufacturing, Topological Design, Mechanical Properties and

Biocompatibility,” Mater. Sci. Eng. C, 76, pp. 1328–1343.

[177] Sing, S. L., An, J., Yeong, W. Y., and Wiria, F. E., 2016, “Laser and Electron-Beam

Powder-Bed Additive Manufacturing of Metallic Implants: A Review on Processes,

Materials and Designs,” J. Orthop. Res., 34(3), pp. 369–385.

[178] Zadpoor, A. A., and Hedayati, R., 2016, “Analytical Relationships for Prediction of the

Mechanical Properties of Additively Manufactured Porous Biomaterials,” J. Biomed.

Mater. Res. - Part A, 104(12), pp. 3164–3174.

[179] Arabnejad, S., Johnston, B., Tanzer, M., and Pasini, D., 2017, “Fully Porous 3D Printed

Titanium Femoral Stem to Reduce Stress-Shielding Following Total Hip Arthroplasty,” J.

Orthop. Res., 35(8), pp. 1774–1783.

[180] De Wild, M., Ghayor, C., Zimmermann, S., Rüegg, J., Nicholls, F., Schuler, F., Chen, T.-

H., and Weber, F. E., 2019, “Osteoconductive Lattice Microarchitecture for Optimized

Bone Regeneration,” 3D Print. Addit. Manuf., 6(1), pp. 40–49.

[181] He, Y., Burkhalter, D., Durocher, D., and Gilbert, J. M., 2018, “Solid-Lattice Hip

Prosthesis Design: Applying Topology and Lattice Optimization to Reduce Stress

Shielding From Hip Implants,” 2018 Design of Medical Devices Conference, American

Society of Mechanical Engineers.

[182] Pronk, T. N., Ayas, C., and Tekõglu, C., 2017, “A Quest for 2D Lattice Materials for

Actuation,” J. Mech. Phys. Solids, 105, pp. 199–216.

[183] Moustafa, A. R., Dinwiddie, R. B., Pawlowski, A. E., Splitter, D. A., Shyam, A., and

Cordero, Z. C., 2018, “Mesostructure and Porosity Effects on the Thermal Conductivity of

Additively Manufactured Interpenetrating Phase Composites,” Addit. Manuf.,

22(February), pp. 223–229.

[184] Takezawa, A., Zhang, X., and Kitamura, M., 2019, “Optimization of an Additively

Manufactured Functionally Graded Lattice Structure with Liquid Cooling Considering

Structural Performances,” Int. J. Heat Mass Transf., 143, p. 118564.

Page 326: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

297

[185] Yu, J., Song, Q. F., Ma, X. L., Zhao, Y. K., Cao, J. H., and Chen, X., 2015, “Study of Heat

Transfer of Composite Lattice Structure for Active Cooling Used in the Scramjet

Combustor,” Mater. Res. Innov., 19, pp. S5843–S5849.

[186] Pawlowski, A. E., Cordero, Z. C., French, M. R., Muth, T. R., Keith Carver, J., Dinwiddie,

R. B., Elliott, A. M., Shyam, A., and Splitter, D. A., 2017, “Damage-Tolerant Metallic

Composites via Melt Infiltration of Additively Manufactured Preforms,” Mater. Des.,

127(March), pp. 346–351.

[187] Brennan-Craddock, J., Brackett, D., Wildman, R., and Hague, R., 2012, “The Design of

Impact Absorbing Structures for Additive Manufacture,” J. Phys. Conf. Ser., 382(1), pp.

2–9.

[188] Ozdemir, Z., Hernandez-Nava, E., Tyas, A., Warren, J. A., Fay, S. D., Goodall, R., Todd,

I., and Askes, H., 2016, “Energy Absorption in Lattice Structures in Dynamics:

Experiments,” Int. J. Impact Eng., 89, pp. 49–61.

[189] Ozdemir, Z., Tyas, A., Goodall, R., and Askes, H., 2017, “Energy Absorption in Lattice

Structures in Dynamics: Nonlinear FE Simulations,” Int. J. Impact Eng., 102.

[190] Hammetter, C. I., Rinaldi, R. G., and Zok, F. W., 2013, “Pyramidal Lattice Structures for

High Strength and Energy Absorption,” J. Appl. Mech., 80(4), p. 041015.

[191] Tancogne-Dejean, T., Spierings, A. B., and Mohr, D., 2016, “Additively-Manufactured

Metallic Micro-Lattice Materials for High Specific Energy Absorption under Static and

Dynamic Loading,” Acta Mater., 116, pp. 14–28.

[192] Helou, M., Vongbunyong, S., and Kara, S., 2016, “Finite Element Analysis and Validation

of Cellular Structures,” Procedia CIRP, 50, pp. 94–99.

[193] Khurana, J., Hanks, B., and Frecker, M., 2018, “Design for Additive Manufacturing of

Cellular Compliant Mechanism Using Thermal History Feedback,” ASME International

Design Engineering Technical Conferences and Comp, pp. 1–15.

[194] Zhang, X. Z., Leary, M., Tang, H. P., Song, T., and Qian, M., 2018, “Selective Electron

Beam Manufactured Ti-6Al-4V Lattice Structures for Orthopedic Implant Applications:

Current Status and Outstanding Challenges,” Curr. Opin. Solid State Mater. Sci., 22(3),

pp. 75–99.

[195] Tao, W., and Leu, M. C., 2016, “Design of Lattice Structure for Additive Manufacturing,”

2016 International Symposium on Flexible Automation (ISFA), IEEE, pp. 325–332.

[196] Wettergreen, M. A., Bucklen, B. S., Starly, B., Yuksel, E., Sun, W., and Liebschner, M. A.

K., 2005, “Creation of a Unit Block Library of Architectures for Use in Assembled

Scaffold Engineering,” CAD Comput. Aided Des., 37(11), pp. 1141–1149.

[197] Satterfield, Z., Kulkarni, N., Fadel, G., Li, G., Coutris, N., and Castanier, M. P., 2017,

“Unit Cell Synthesis for Design of Materials With Targeted Nonlinear Deformation

Response,” J. Mech. Des., 139(12), p. 121401.

Page 327: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

298

[198] Fazelpour, M., Summers, J. D., and Bluoin, V. Y., 2016, “A Taxonomy For Representing

Prismatic Cellular Materials,” ASME 2016 Int. Des. Eng. Tech. Conf. Comput. Inf. Eng.

Conf., 2, p. V02BT03A001.

[199] Zok, F. W., Latture, R. M., and Begley, M. R., 2016, “Periodic Truss Structures,” J. Mech.

Phys. Solids, 96, pp. 184–203.

[200] Bhate, D., Penick, C., Ferry, L., and Lee, C., 2018, “Classification and Selection of

Cellular Materials in Mechanical Design : Engineering and Biomimetic Approaches,” pp.

1–27.

[201] Deshpande, V. S., Fleck, N. A., and Ashby, M. F., 2001, “Effective Properties of the

Octet-Truss Lattice Material,” J. Mech. Phys. Solids, 49(8), pp. 1747–1769.

[202] Hedayati, R., Sadighi, M., Mohammadi-Aghdam, M., and Zadpoor, A. A., 2016, “Effect

of Mass Multiple Counting on the Elastic Properties of Open-Cell Regular Porous

Biomaterials,” Mater. Des., 89, pp. 9–20.

[203] Hedayati, R., Sadighi, M., Mohammadi-Aghdam, M., and Zadpoor, A. A., 2016,

“Mechanical Properties of Regular Porous Biomaterials Made from Truncated Cube

Repeating Unit Cells: Analytical Solutions and Computational Models,” Mater. Sci. Eng.

C, 60, pp. 163–183.

[204] Hedayati, R., Sadighi, M., Mohammadi-Aghdam, M., and Zadpoor, A. A., 2016,

“Mechanics of Additively Manufactured Porous Biomaterials Based on the

Rhombicuboctahedron Unit Cell,” J. Mech. Behav. Biomed. Mater., 53, pp. 272–294.

[205] Hedayati, R., Sadighi, M., Mohammadi-Aghdam, M., and Zadpoor, A. A., 2017,

“Analytical Relationships for the Mechanical Properties of Additively Manufactured

Porous Biomaterials Based on Octahedral Unit Cells,” Appl. Math. Model., 46, pp. 408–

422.

[206] Hedayati, R., Sadighi, M., Mohammadi-Aghdam, M., and Zadpoor, A. A., 2016,

“Mechanical Behavior of Additively Manufactured Porous Biomaterials Made from

Truncated Cuboctahedron Unit Cells,” Int. J. Mech. Sci., 106, pp. 19–38.

[207] Lei, H., Li, C., Meng, J., Zhou, H., Liu, Y., Zhang, X., Wang, P., and Fang, D., 2019,

“Evaluation of Compressive Properties of SLM-Fabricated Multi-Layer Lattice Structures

by Experimental Test and μ-CT-Based Finite Element Analysis,” Mater. Des., 169, p.

107685.

[208] Han, C., Yan, C., Wen, S., Xu, T., Li, S., Liu, J., Wei, Q., and Shi, Y., 2017, “Effects of

the Unit Cell Topology on the Compression Properties of Porous Co-Cr Scaffolds

Fabricated via Selective Laser Melting,” Rapid Prototyp. J., 23(1), pp. 16–27.

[209] Ahmadi, S. M., Campoli, G., Amin Yavari, S., Sajadi, B., Wauthle, R., Schrooten, J.,

Weinans, H., and Zadpoor, A. A., 2014, “Mechanical Behavior of Regular Open-Cell

Porous Biomaterials Made of Diamond Lattice Unit Cells,” J. Mech. Behav. Biomed.

Mater., 34, pp. 106–115.

Page 328: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

299

[210] Tang, T. L. E., Liu, Y., Lu, D., Arisoy, E. B., and Musuvathy, S., 2017, “Lattice Structure

Design Advisor for Additive Manufacturing Using Gaussian Process,” ASME

International Design Engineering Technical Conferences and Computers and Information

in Engineering Conference, pp. 1–10.

[211] Ahmadi, S. M., Yavari, S. A., Wauthle, R., Pouran, B., Schrooten, J., Weinans, H., and

Zadpoor, A. A., 2015, “Additively Manufactured Open-Cell Porous Biomaterials Made

from Six Different Space-Filling Unit Cells: The Mechanical and Morphological

Properties,” Materials (Basel)., 8(4), pp. 1871–1896.

[212] Cheng, X. Y., Li, S. J., Murr, L. E., Zhang, Z. B., Hao, Y. L., Yang, R., Medina, F., and

Wicker, R. B., 2012, “Compression Deformation Behavior of Ti-6Al-4V Alloy with

Cellular Structures Fabricated by Electron Beam Melting,” J. Mech. Behav. Biomed.

Mater., 16(1), pp. 153–162.

[213] Aremu, A. O., Maskery, I., Tuck, C., Ashcroft, I. A., Wildman, R. D., and Hague, R. I. .,

2014, “A Comparative Finite Element Study of Cubic Unit Cells for Selective Laser

Melting,” Int. Solid Free. Fabr. Symp., pp. 1238–1249.

[214] Smith, M., Guan, Z., and Cantwell, W. J., 2013, “Finite Element Modelling of the

Compressive Response of Lattice Structures Manufactured Using the Selective Laser

Melting Technique,” Int. J. Mech. Sci., 67, pp. 28–41.

[215] Ushijima, K., Cantwell, W. J., and Chen, D. H., 2013, “Prediction of the Mechanical

Properties of Micro-Lattice Structures Subjected to Multi-Axial Loading,” Int. J. Mech.

Sci., 68, pp. 47–55.

[216] Ushijima, K., Cantwell, W. J., Mines, R. A. W., Tsopanos, S., and Smith, M., 2011, “An

Investigation into the Compressive Properties of Stainless Steel Micro-Lattice Structures,”

J. Sandw. Struct. Mater., 13(3), pp. 303–329.

[217] Brodin, H., and Saarimäki, J., 2013, “Mechanical Properties of Lattice Truss Structures

Made of a Selective Laser Melted Superalloy,” 13th International Conference on Fracture

(ICF13), pp. 1–10.

[218] Yan, C., Hao, L., Hussein, A., Bubb, S. L., Young, P., and Raymont, D., 2014,

“Evaluation of Light-Weight AlSi10Mg Periodic Cellular Lattice Structures Fabricated via

Direct Metal Laser Sintering,” J. Mater. Process. Technol., 214(4), pp. 856–864.

[219] Park, S. I., Rosen, D. W., Choi, S. kyum, and Duty, C. E., 2014, “Effective Mechanical

Properties of Lattice Material Fabricated by Material Extrusion Additive Manufacturing,”

Addit. Manuf., 1, pp. 12–23.

[220] Hussein, A. Y., 2013, “The Development of Lightweight Cellular Structures for Metal

Additive Manufacturing,” University of Exeter.

[221] Yan, C., Hao, L., Hussein, A., Young, P., Huang, J., and Zhu, W., 2015, “Microstructure

and Mechanical Properties of Aluminium Alloy Cellular Lattice Structures Manufactured

by Direct Metal Laser Sintering,” Mater. Sci. Eng. A, 628, pp. 238–246.

Page 329: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

300

[222] Yan, C., Hao, L., Hussein, A., and Raymont, D., 2012, “Evaluations of Cellular Lattice

Structures Manufactured Using Selective Laser Melting,” Int. J. Mach. Tools Manuf., 62,

pp. 32–38.

[223] Yan, C., Hao, L., Hussein, A., Young, P., and Raymont, D., 2014, “Advanced Lightweight

316L Stainless Steel Cellular Lattice Structures Fabricated via Selective Laser Melting,”

Mater. Des., 55, pp. 533–541.

[224] Nguyen, J., Park, S., Rosen, D. W., Folgar, L., and Williams, J., 2012, “Conformal Lattice

Structure Design and Fabrication,” Solid Freeform Fabrication Symposium, pp. 138–161.

[225] Wang, H., Chen, Y., and Rosen, D. W., 2005, “A Hybrid Geometric Modeling Method for

Large Scale Conformal Cellular Structures,” 25th Comput. Inf. Eng. Conf. Parts A B, 3,

pp. 421–427.

[226] Wang, H. V., and Rosen, D. W., 2002, “Computer-Aided Design Methods for the

Additive Fabrication of Truss Structures,” Int. Conf. Manuf. Autom., (August), pp. 181–

198.

[227] Engelbrecht, S., Luis Folgar, Rosen, D. W., Schulberger, G., and Williams, J., 2009,

“Cellular Structures for Optimal Performance,” Proc. SFF …, pp. 831–842.

[228] Bhate, D., 2019, “Four Questions in Cellular Material Design,” Materials (Basel)., 12(7).

[229] Brackett, D., Ashcroft, I., and Hague, R., 2011, “A Dithering Based Method to Generate

Variable Volume Lattice Cells for Additive Manufacturing,” 22nd Annual International

Solid Freeform Fabrication Symposium, pp. 671–679.

[230] Arabnejad Khanoki, S., and Pasini, D., 2012, “Multiscale Design and Multiobjective

Optimization of Orthopedic Hip Implants with Functionally Graded Cellular Material,” J.

Biomech. Eng., 134(3), p. 031004.

[231] Arabnejad Khanoki, S., and Pasini, D., 2013, “The Fatigue Design of a Bone Preserving

Hip Implant With Functionally Graded Cellular Material,” J. Med. Device., 7(2), pp. 1–2.

[232] Arabnejad Khanoki, S., and Pasini, D., 2013, “Fatigue Design of a Mechanically

Biocompatible Lattice for a Proof-of-Concept Femoral Stem,” J. Mech. Behav. Biomed.

Mater., 22, pp. 65–83.

[233] Graf, G. C., Chu, J., Engelbrecht, S., and Rosen, D. W., 2009, “Synthesis Methods for

Lightweight Lattice Structures,” ASME International Design Engineering Technical

Conferences and Computers and Information in Engineering Conference, pp. 1–11.

[234] Chang, P. S., and Rosen, D. W., 2013, “The Size Matching and Scaling Method: A

Synthesis Method for the Design of Mesoscale Cellular Structures,” Int. J. Comput. Integr.

Manuf., 26(10), pp. 907–927.

[235] nTopology, “NTop Platform” [Online]. Available: https://ntopology.com/.

[236] Paramatters Inc, “Paramatters” [Online]. Available: https://paramatters.com/.

Page 330: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

301

[237] 3D Systems, “3DXpert” [Online]. Available:

https://www.3dsystems.com/software/3dxpert.

[238] ANSYS Inc, 2020, “Discovery SpaceClaim” [Online]. Available:

https://www.ansys.com/products/3d-design/ansys-spaceclaim.

[239] Naing, M. W., Chua, C. K., Leong, K. F., and Wang, Y., 2005, “Fabrication of

Customised Scaffolds Using Computer‐aided Design and Rapid Prototyping Techniques,”

Rapid Prototyp. J., 11(4), pp. 249–259.

[240] Cheah, C., Chua, C., Leong, K., Cheong, C.-H., and Naing, M.-W., 2004, “Automatic

Algorithm for Generating Complex Polyhedral Scaffold Structures for Tissue

Engineering,” Tissue Eng., 10(3–4), pp. 595–610.

[241] Sudarmadji, N., Tan, J. Y., Leong, K. F., Chua, C. K., and Loh, Y. T., 2011,

“Investigation of the Mechanical Properties and Porosity Relationships in Selective Laser-

Sintered Polyhedral for Functionally Graded Scaffolds,” Acta Biomater., 7(2), pp. 530–

537.

[242] Vongbunyong, S., and Kara, S., 2016, “Development of Software Tool for Cellular

Structure Integration for Additive Manufacturing,” Procedia CIRP, 48, pp. 489–494.

[243] Vongbunyong, S., and Kara, S., 2017, “Rapid Generation of Uniform Cellular Structure

by Using Prefabricated Unit Cells,” Int. J. Comput. Integr. Manuf., 30(8), pp. 792–804.

[244] Nazir, A., Abate, K. M., Kumar, A., and Jeng, J. Y., 2019, “A State-of-the-Art Review on

Types, Design, Optimization, and Additive Manufacturing of Cellular Structures,” Int. J.

Adv. Manuf. Technol., 104(9–12), pp. 3489–3510.

[245] Koizumi, Y., Okazaki, A., Chiba, A., Kato, T., and Takezawa, A., 2016, “Cellular Lattices

of Biomedical Co-Cr-Mo-Alloy Fabricated by Electron Beam Melting with the Aid of

Shape Optimization,” Addit. Manuf., 12, pp. 305–313.

[246] Savio, G., Meneghello, R., and Concheri, G., 2018, “Geometric Modeling of Lattice

Structures for Additive Manufacturing,” Rapid Prototyp. J., 24(2), pp. 351–360.

[247] McMillan, M. L., Jurg, M., Leary, M., and Brandt, M., 2017, “Programmatic Generation

of Computationally Efficient Lattice Structures for Additive Manufacture,” Rapid

Prototyp. J., 23(3), pp. 486–494.

[248] Gupta, A., Kurzeja, K., Rossignac, J., Allen, G., Kumar, P. S., and Musuvathy, S., 2019,

“Programmed-Lattice Editor and Accelerated Processing of Parametric Program-

Representations of Steady Lattices,” CAD Comput. Aided Des., 113, pp. 35–47.

[249] Goel, A., and Anand, S., 2019, “Design of Functionally Graded Lattice Structures Using

B-Splines for Additive Manufacturing,” Procedia Manuf., 34, pp. 655–665.

[250] Maconachie, T., Leary, M., Lozanovski, B., Zhang, X., Qian, M., Faruque, O., and Brandt,

M., 2019, “SLM Lattice Structures: Properties, Performance, Applications and

Challenges,” Mater. Des., p. 108137.

Page 331: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

302

[251] Hanzl, P., Zetkova, I., and Dana, M., 2017, “Issues of Lattice Structures Production via

Metal Additive Manufacturing,” Manuf. Technol., 17(6).

[252] Tang, Y., Dong, G., Zhou, Q., and Zhao, Y. F., 2018, “Lattice Structure Design and

Optimization with Additive Manufacturing Constraints,” IEEE Trans. Autom. Sci. Eng.,

15(4), pp. 1546–1562.

[253] Behera, S., Sahoo, S., and Pati, B. B., 2015, “A Review on Optimization Algorithms and

Application to Wind Energy Integration to Grid,” Renew. Sustain. Energy Rev., 48, pp.

214–227.

[254] Venter, G., 2010, “Review of Optimization Techniques,” Encycl. Aerosp. Eng., pp. 1–12.

[255] Marthaler, D. E., 2013, “An Overview of Mathematical Methods for Numerical

Optimization,” Numerical Methods for Metamaterial Design, pp. 31–54.

[256] Vikhar, P. A., 2016, “Evolutionary Algorithms : A Critical Review and Its Future

Prospects,” pp. 261–265.

[257] Deb, K., Pratap, A., Agarwal, S., and Meyarivan, T., 2002, “A Fast and Elitist

Multiobjective Genetic Algorithm: NSGA-II,” IEEE Trans. Evol. Comput., 6(2), pp. 182–

197.

[258] Hadka, D., and Reed, P., 2013, “Borg: An Auto-Adaptive Many-Objective Evolutionary

Computing Framework,” Evol. Comput., 21(2), pp. 231–259.

[259] Kim, S., 2006, “Gradient-Based Simulation Optimization,” IEEE Winter Simulation

Conference, pp. 159–167.

[260] Ruder, S., 2016, “An Overview of Gradient Descent Optimization Algorithms,” CoRR2,

arXiv:1609.

[261] Golfetto, W. A., and Fernandes, S. da S., 2012, “A Review of Gradient Algorithms for

Numerical Computation of Optimal Trajectories,” J. Aerosp. Technol. Manag., 4(2), pp.

131–143.

[262] Salomon, R., 1998, “Evolutionary Algorithms and Gradient Search: Similarities and

Differences,” IEEE Trans. Evol. Comput., 2(2), pp. 45–55.

[263] Zingg, D. W., Nemec, M., and Pulliam, T. H., 2008, “A Comparative Evaluation of

Genetic and Gradient-Based Algorithms Applied to Aerodynamic Optimization,” Rev.

Eur. mécanique numérique, 17, pp. 103–126.

[264] Haupt, R., 1995, “Comparison between Genetic and Gradient-Based Optimization

Algorithms for Solving Electromagnetics Problems,” IEEE Trans. Magn., 31(3), pp.

1932–1935.

[265] Ahmad, F., Isa, N. A. M., Osman, M. K., and Hussain, Z., 2010, “Performance

Comparison of Gradient Descent and Genetic Algorithm Based Artificial Neural

Networks Training,” IEEE Intelligent Systems Design and Applications, ISDA, Cairo, pp.

Page 332: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

303

604–609.

[266] Song, Y., and Xi, P., 2009, “Genetic Algorithm and Gradient-Based Algorithm

Optimization of Vehicle Turning Mechanism,” IEEE International Conference on

Computer Science and Information Technology, ICCSIT, pp. 81–84.

[267] Montastruc, L., Azzaro-Pantel, C., Pibouleau, L., and Domenech, S., 2004, “Use of

Genetic Algorithms and Gradient Based Optimization Techniques for Calcium Phosphate

Precipitation,” Chem. Eng. Process. Process Intensif., 43(10), pp. 1289–1298.

[268] Bendsøe, M. P., and Kikuchi, N., 1988, “Generating Optimal Topologies in Structural

Design Using a Homogenization Method,” Comput. Methods Appl. Mech. Eng., 71(2),

pp. 197–224.

[269] Bendsøe, M. P., 1989, “Optimal Shape Design as a Material Distribution Problem,” Struct.

Optim., 1(4), pp. 193–202.

[270] Bendsøe, M. P., and Sigmund, O., 1999, “Material Interpolation Schemes in Topology

Optimization,” Arch. Appl. Mech., 69(9–10), pp. 635–654.

[271] Frecker, M. I., Ananthasuresh, G. K., Njshiwaki, S., Kikuchi, N., and Kota, S., 1997,

“Topological Synthesis of Compliant Mechanisms Using Multi-Criteria Optimization,” J.

Mech. Des., 119(June).

[272] Yin, L., and Ananthasuresh, G. K., 2001, “Topology Optimization of Compliant

Mechanisms with Multiple Materials Using a Peak Function Material Interpolation

Scheme,” Struct. Multidiscip. Optim., 23(1), pp. 49–62.

[273] Guest, J. K., Asadpoure, A., and Ha, S. H., 2011, “Eliminating Beta-Continuation from

Heaviside Projection and Density Filter Algorithms,” Struct. Multidiscip. Optim., 44(4),

pp. 443–453.

[274] Mankame, N. D., and Ananthasuresh, G. K., 2004, “Topology Optimization for Synthesis

of Contact-Aided Compliant Mechanisms Using Regularized Contact Modeling,”

Comput. Struct., 82(15–16), pp. 1267–1290.

[275] De Leon, D. M., Alexandersen, J., Jun, J. S., and Sigmund, O., 2015, “Stress-Constrained

Topology Optimization for Compliant Mechanism Design,” Struct. Multidiscip. Optim.,

52(5), pp. 929–943.

[276] Frecker, M., Kikuchi, N., and Kota, S., 1999, “Topology Optimization of Compliant

Mechanisms with Multiple Outputs,” Struct. Optim., 17(4), pp. 269–278.

[277] Santos Iwamura, R., and Rocha De Faria, A., 2013, “Topology Optimization of Multiple

Load Case Structures,” International Symposium on Solid Mechanics, Porto Alegre, pp. 1–

11.

[278] Lohan, D. J., Dede, E. M., and Allison, J. T., 2017, “Topology Optimization for Heat

Conduction Using Generative Design Algorithms,” Struct. Multidiscip. Optim., 55(3), pp.

1063–1077.

Page 333: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

304

[279] Wang, N., and Zhang, X., 2014, “Multi-Material Topology Optimization of Compliant

Mechanism Using Ground Structure Approach,” 2014 Int. Conf. Manip. Manuf. Meas.

Nanoscale, (October), pp. 249–254.

[280] Liu, J., and Ma, Y., 2016, “A Survey of Manufacturing Oriented Topology Optimization

Methods,” Adv. Eng. Softw., 100, pp. 161–175.

[281] Bendsøe, M. P., and Sigmund, O., 2004, Topology Optimization, Springer Berlin

Heidelberg, Berlin, Heidelberg.

[282] Sigmund, O., and Maute, K., 2013, “Topology Optimization Approaches: A Comparative

Review,” Struct. Multidiscip. Optim., 48(6), pp. 1031–1055.

[283] Reddy K., S. N., Ferguson, I., Frecker, M., and Simpson, T. W., 2016, “Topology

Optimization Software for Additive Manufacturing: A Review of Current Capabilities and

a Real-World Example,” Proc. ASME 2016 Int. Des. Eng. Tech. Conf. Comput. Inf. Eng.

Conf.

[284] Ananthasuresh, G. K., and Frecker, M. I., 2001, “Optimal Synthesis with Continuum

Models,” Compliant Mechanisms, L.L. Howell, ed., John Wiley & Sons.

[285] Huang, X., Li, Y., Zhou, S. W., and Xie, Y. M., 2014, “Topology Optimization of

Compliant Mechanisms with Desired Structural Stiffness,” Eng. Struct., 79(May 2018),

pp. 13–21.

[286] Dorn, W. S., Gomory, R. E., and Greenberg, H. G., 1964, “Automatic Design of Optimal

Structures,” J. Mec., 3(1).

[287] Ohsaki, M., 2016, Optimization of Finite Dimensional Structures, CRC Press.

[288] Zhang, X. S., Paulino, G. H., and Ramos, A. S., 2018, “Multi-Material Topology

Optimization with Multiple Volume Constraints: A General Approach Applied to Ground

Structures with Material Nonlinearity,” Struct. Multidiscip. Optim., 57(1), pp. 161–182.

[289] Liu, K., Paulino, G. H., and Gardoni, P., 2016, “Reliability-Based Topology Optimization

Using a New Method for Sensitivity Approximation - Application to Ground Structures,”

Struct. Multidiscip. Optim., 54(3), pp. 553–571.

[290] Zegard, T., and Paulino, G. H., 2014, “GRAND — Ground Structure Based Topology

Optimization for Arbitrary 2D Domains Using MATLAB,” Struct. Multidiscip. Optim.,

50(5), pp. 861–882.

[291] Zegard, T., and Paulino, G. H., 2015, “GRAND3 — Ground Structure Based Topology

Optimization for Arbitrary 3D Domains Using MATLAB,” Struct. Multidiscip. Optim.,

52(6), pp. 1161–1184.

[292] Hagishita, T., and Ohsaki, M., 2009, “Topology Optimization of Trusses by Growing

Ground Structure Method,” Struct. Multidiscip. Optim., 37(4), pp. 377–393.

[293] Cui, H., An, H., and Huang, H., 2018, “Truss Topology Optimization Considering Local

Page 334: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

305

Buckling Constraints and Restrictions on Intersection and Overlap of Bar Members,”

Struct. Multidiscip. Optim., pp. 1–20.

[294] Asadpoure, A., Guest, J. K., and Valdevit, L., 2015, “Incorporating Fabrication Cost into

Topology Optimization of Discrete Structures and Lattices,” Struct. Multidiscip. Optim.,

51(2), pp. 385–396.

[295] Liu, J., Gaynor, A. T., Chen, S., Kang, Z., Suresh, K., Takezawa, A., Li, L., Kato, J.,

Tang, J., Wang, C. C. L., Cheng, L., Liang, X., and To, A. C., 2018, “Current and Future

Trends in Topology Optimization for Additive Manufacturing,” Struct. Multidiscip.

Optim., pp. 2457–2483.

[296] Zegard, T., and Paulino, G. H., 2016, “Bridging Topology Optimization and Additive

Manufacturing,” Struct. Multidiscip. Optim., 53, pp. 175–192.

[297] Sigmund, O., 1995, “Tailoring Materials with Prescribed Elastic Properties,” Mech.

Mater., 20(4), pp. 351–368.

[298] Sigmund, O., 1994, “Materials With Prescribed Constitutive Parameters: An Inverse

Homogenization Problem,” Int. J. Solids Struct., 31(17), pp. 2313–2329.

[299] Guth, D. C., Luersen, M. A., and Muñoz-Rojas, P. A., 2012, “Optimization of Periodic

Truss Materials Including Constitutive Symmetry Constraints,” Materwiss. Werksttech.,

43(5), pp. 447–456.

[300] Guth, D. C., Luersen, M. A., and Muñoz-Rojas, P. A., 2015, “Optimization of Three-

Dimensional Truss-like Periodic Materials Considering Isotropy Constraints,” Struct.

Multidiscip. Optim., 52(5), pp. 889–901.

[301] Messner, M. C., 2016, “Optimal Lattice-Structured Materials,” J. Mech. Phys. Solids, 96,

pp. 162–183.

[302] Neves, M. M., Sigmund, O., and Bendsøe, M. P., 2002, “Topology Optimization of

Periodic Microstructures with a Penalization of Highly Localized Buckling Modes,” Int. J.

Numer. Methods Eng., 54(6), pp. 809–834.

[303] Liu, L., Yan, J., and Cheng, G., 2008, “Optimum Structure with Homogeneous Optimum

Truss-like Material,” Comput. Struct., 86(13–14), pp. 1417–1425.

[304] Huang, X., Zhou, S. W., Xie, Y. M., and Li, Q., 2013, “Topology Optimization of

Microstructures of Cellular Materials and Composites for Macrostructures,” Comput.

Mater. Sci., 67, pp. 397–407.

[305] De Pasquale, G., Montemurro, M., Catapano, A., Bertolino, G., and Revelli, L., 2018,

“Cellular Structures from Additive Processes: Design, Homogenization and Experimental

Validation,” Procedia Struct. Integr., 8, pp. 75–82.

[306] Cheng, G. D., Cai, Y. W., and Xu, L., 2013, “Novel Implementation of Homogenization

Method to Predict Effective Properties of Periodic Materials,” Acta Mech. Sin. Xuebao,

29(4), pp. 550–556.

Page 335: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

306

[307] Park, S., and Rosen, D. W., 2016, “Homogenization of Mechanical Properties for

Additively,” ASME International Design Engineering Technical Conferences, pp. 1–10.

[308] Arabnejad, S., and Pasini, D., 2013, “Mechanical Properties of Lattice Materials via

Asymptotic Homogenization and Comparison with Alternative Homogenization

Methods,” Int. J. Mech. Sci., 77, pp. 249–262.

[309] Vigliotti, A., and Pasini, D., 2012, “Stiffness and Strength of Tridimensional Periodic

Lattices,” Comput. Methods Appl. Mech. Eng., 229–232, pp. 27–43.

[310] Xia, L., and Breitkopf, P., 2015, “Design of Materials Using Topology Optimization and

Energy-Based Homogenization Approach in Matlab,” Struct. Multidiscip. Optim., 52(6),

pp. 1229–1241.

[311] Neves, M. M., Rodrigues, H., and Guedes, J. M., 2000, “Optimal Design of Periodic

Linear Elastic Microstructures,” Comput. Struct., 76.

[312] Seepersad, C. C., Allen, J. K., McDowell, D. L., and Mistree, F., 2006, “Robust Design of

Cellular Materials With Topological and Dimensional Imperfections,” J. Mech. Des.,

128(6), p. 1285.

[313] Watts, S., Arrighi, W., Kudo, J., Tortorelli, D. A., and White, D. A., 2019, “Simple,

Accurate Surrogate Models of the Elastic Response of Three-Dimensional Open Truss

Micro-Architectures with Applications to Multiscale Topology Design,” Struct.

Multidiscip. Optim., 3, pp. 1887–1920.

[314] Watts, S., and Tortorelli, D. A., 2017, “A Geometric Projection Method for Designing

Three-Dimensional Open Lattices with Inverse Homogenization,” Int. J. Numer. Methods

Eng., 112(11), pp. 1564–1588.

[315] Norato, J. A., Bell, B. K., and Tortorelli, D. A., 2015, “A Geometry Projection Method for

Continuum-Based Topology Optimization with Discrete Elements,” Comput. Methods

Appl. Mech. Eng., 293, pp. 306–327.

[316] Sharpe, C., Seepersad, C. C., Watts, S., and Tortorelli, D., 2018, “Design of Mechanical

Metamaterials via Constrained Bayesian Optimization,” Volume 2A: 44th Design

Automation Conference, American Society of Mechanical Engineers, pp. 1–11.

[317] Zhang, S., Norato, J. A., Gain, A. L., and Lyu, N., 2016, “A Geometry Projection Method

for the Topology Optimization of Plate Structures,” Struct. Multidiscip. Optim., 54(5), pp.

1173–1190.

[318] Norato, J. A., 2018, “Topology Optimization with Supershapes,” Struct. Multidiscip.

Optim., 58(2), pp. 415–434.

[319] Wang, Y., Zhang, L., Daynes, S., Zhang, H., Feih, S., and Wang, M. Y., 2018, “Design of

Graded Lattice Structure with Optimized Mesostructures for Additive Manufacturing,”

Mater. Des., 142, pp. 114–123.

[320] Cheng, L., Zhang, P., Biyikli, E., Bai, J., Robbins, J., and To, A., 2017, “Efficient Design

Page 336: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

307

Optimization of Variable-Density Cellular Structures for Additive Manufacturing: Theory

and Experimental Validation,” Rapid Prototyp. J., 23(4), pp. 660–677.

[321] Han, Y., and Lu, W. F., 2018, “Optimization Design of Nonuniform Cellular Structures

for Additive Manufacturing,” Vol. 1 Addit. Manuf. Bio Sustain. Manuf., p.

V001T01A033.

[322] Panesar, A., Abdi, M., Hickman, D., and Ashcroft, I., 2018, “Strategies for Functionally

Graded Lattice Structures Derived Using Topology Optimisation for Additive

Manufacturing,” Addit. Manuf., 19, pp. 81–94.

[323] Kang, D., Park, S., Son, Y., Yeon, S., Kim, S. H., and Kim, I., 2019, “Multi-Lattice Inner

Structures for High-Strength and Light-Weight in Metal Selective Laser Melting Process,”

Mater. Des., 175, p. 107786.

[324] Steuben, J. C., Iliopoulos, A. P., and Michopoulos, J. G., 2018, “Multiscale Topology

Optimization for Additively Manufactured Objects,” J. Comput. Inf. Sci. Eng., 18(3), pp.

1–9.

[325] Sabiston, G., Ryan, L., and Kim, I. Y., 2018, “Minimization of Cost and Print Time of

Additive Manufacturing Via Topology,” ASME International Design Engineering

Technical Conferences & Computers and Information in Engineering, pp. 1–10.

[326] Langelaar, M., 2017, “An Additive Manufacturing Filter for Topology Optimization of

Print-Ready Designs,” Struct. Multidiscip. Optim., 55(3), pp. 871–883.

[327] Langelaar, M., 2016, “Topology Optimization of 3D Self-Supporting Structures for

Additive Manufacturing,” Addit. Manuf., 12, pp. 60–70.

[328] Langelaar, M., 2016, “Topology Optimization for Additive Manufacturing With

Controllable Support Structures,” ECCOMAS Congress.

[329] Brackett, D., Ashcroft, I., and Hague, R., 2011, “Topology Optimization for Additive

Manufacturing,” Solid Freeform Fabrication Symposium, pp. 348–362.

[330] Gaynor, A. T., and Guest, J. K., 2016, “Topology Optimization Considering Overhang

Constraints: Eliminating Sacrificial Support Material in Additive Manufacturing through

Design,” Struct. Multidiscip. Optim., 54(5), pp. 1157–1172.

[331] Liu, S., Li, Q., Chen, W., Tong, L., and Cheng, G., 2015, “An Identification Method for

Enclosed Voids Restriction in Manufacturability Design for Additive Manufacturing

Structures,” Front. Mech. Eng., 10(2), pp. 126–137.

[332] Schwerdtfeger, J., Wein, F., Leugering, G., Singer, R. F., Körner, C., Stingl, M., and

Schury, F., 2011, “Design of Auxetic Structures via Mathematical Optimization,” Adv.

Mater., 23(22–23), pp. 2650–2654.

[333] Lin, C.-Y., Wirtz, T., LaMarca, F., and Hollister, S. J., 2015, “Structural and Mechanical

Evaluations of a Topology Optimized Titanium Interbody Fusion Cage Fabricated by

Selective Laser Melting Process,” Clin. Exp. Rheumatol., 33, pp. 97–103.

Page 337: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

308

[334] Tamburrino, F., Graziosi, S., and Bordegoni, M., 2018, “The Design Process of Additively

Manufactured Mesoscale Lattice Structures: A Review,” J. Comput. Inf. Sci. Eng., 18(4),

pp. 1–16.

[335] Meng, L., Zhang, W., Quan, D., Shi, G., Tang, L., Hou, Y., Breitkopf, P., Zhu, J., and

Gao, T., 2019, “From Topology Optimization Design to Additive Manufacturing: Today’s

Success and Tomorrow’s Roadmap,” Arch. Comput. Methods Eng., (0123456789).

[336] Plocher, J., and Panesar, A., 2019, “Review on Design and Structural Optimisation in

Additive Manufacturing: Towards next-Generation Lightweight Structures,” Mater. Des.,

183, p. 108164.

[337] Bacciaglia, A., Ceruti, A., and Liverani, A., 2020, “Additive Manufacturing Challenges

and Future Developments in the Next Ten Years,” Design Tools and Methods in Industrial

Engineering, Springer International Publishing, pp. 891–902.

[338] Hanks, B., Frecker, M., and Moyer, M., 2016, “Design of a Compliant Endoscopic

Ultrasound-Guided Radiofrequency Ablation Probe,” ASME International Design

Engineering Technical Conferences and Computers and Information in Engineering

Conference, pp. 1–9.

[339] Hanks, B., Frecker, M., and Moyer, M., 2018, “Optimization of an Endoscopic

Radiofrequency Ablation Electrode,” ASME J. Med. Devices, 12(3), pp. 1–11.

[340] Misra, S., Reed, K. B., Schafer, B. W., Ramesh, K. T., and Okamura, A. M., 2010,

“Mechanics of Flexible Needles Robotically Steered through Soft Tissue,” Int. J. Rob.

Res., 29(13), pp. 1640–1660.

[341] Webster III, R. J., Memisevic, J., and Okamura, a M., 2005, “Design Considerations for

Robotic Needle Steering,” Proc. IEEE Int. Conf. Robot. Autom., (April), pp. 3599–3605.

[342] Webster III, R. J., Cowan, N. J., Chirikjian, G. S., and Okamura, a M., 2006,

“Nonholonomic Modelling of Needle Steering,” Int. J. Rob. Res., 25(5/6), pp. 509–526.

[343] COMSOL Inc, “COMSOL Multiphysics” [Online]. Available: https://www.comsol.com/.

[344] MathWorks, “MATLAB” [Online]. Available:

https://www.mathworks.com/products/matlab.html.

[345] Hanks, B., Frecker, M., and Moyer, M., 2017, “Optimization of a Compliant Endoscopic

Radiofrequency Ablation Electrode,” ASME International Design Engineering Technical

Conferences and Computers and Information in Engineering Conference, DETC2017-

673576, Cleveland.

[346] Mathworks, 2016, Global Optimization Toolbox.

[347] Hasgall, P., Di Gennaro, F., Baumgartner, C., Neufeld, E., Gosselin, M., Payne, D.,

Klingenbock, A., and Kuster, N., 2015, “IT’IS Database for Thermal and Electromagnetic

Parameters of Biological Tissues,” Version 3.0 [Online]. Available:

www.itis.ethz.ch/database.

Page 338: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

309

[348] Gabriel, C., 1996, Compilation of the Dielectric Properties of the Body Tissues at RF and

Microwave Frequencies, London.

[349] Komar, G., Kauhanen, S., Liukko, K., Seppänen, M., Kajander, S., Ovaska, J., Nuutila, P.,

and Minn, H., 2009, “Decreased Blood Flow with Increased Metabolic Activity: A Novel

Sign of Pancreatic Tumor Aggressiveness,” Clin. Cancer Res., 15(17), pp. 5511–5517.

[350] Bergman, T. L., Lavine, A. S., Incropera, F. P., and Dewitt, D. P., 2011, Fundamentals of

Heat and Mass Transfer, Wiley.

[351] Committee, A. I. H., 1995, “Electrical Testing of Polymers,” Engineered Materials

Handbook Desk Edition, ASM International, Materials Park, Ohio, pp. 423–438.

[352] 2015, “Nitinol Technical Properties,” Johnson Matthey Med. Components, pp. 1–2

[Online]. Available: http://jmmedical.com/resources/221/Nitinol-Technical-

Properties.html.

[353] Rossman, C., and Haemmerich, D., 2014, “Review of Temperature Dependence of

Thermal Properties, Dielectric Properties, and Perfusion of Biological Tissues at

Hyperthermic and Ablation Temperatures,” Crit Rev Biomed Eng, 42(6), pp. 467–492.

[354] Rasband, W. S. W., 2012, “ImageJ,” U. S. Natl. Institutes Heal. Bethesda, Maryland,

USA.

[355] Hanks, B., Azhar, F., Frecker, M., Clement, R., Greaser, J., and Snook, K., 2019, “A

Deployable Multi-Tine Endoscopic Radiofrequency Ablation Electrode: Simulation

Validation in a Thermochromic Tissue Phantom,” Design of Medical Devices Conference,

DMD2019-3214.

[356] Datla, N. V., Konh, B., Koo, J. J. Y., Choi, D. J. W., Yu, Y., Dicker, A. P., Podder, T. K.,

Darvish, K., and Hutapea, P., 2014, “Polyacrylamide Phantom for Self-Actuating Needle-

Tissue Interaction Studies,” Med. Eng. Phys., 36(1), pp. 140–145.

[357] Negussie, A. H., Partanen, A., Mikhail, A. S., Xu, S., Abi-Jaoudeh, N., Maruvada, S., and

Wood, B. J., 2016, “Thermochromic Tissue-Mimicking Phantom for Optimisation of

Thermal Tumour Ablation,” Int. J. Hyperth., 32(3), pp. 239–243.

[358] Mathworks, “Fmincon” [Online]. Available:

https://www.mathworks.com/help/optim/ug/fmincon.html.

[359] Deb, K., Mohan, M., and Mishra, S., 2003, A Fast Multi-Objective Evolutionary

Algorithm for Finding Well-Spread Pareto-Optimal Solutions.

[360] van der Sluis, O., Schreurs, P. J. G., Brekelmans, W. A. M., and Meijer, H. E. H., 2000,

“Overall Behaviour of Heterogeneous Elastoviscoplastic Materials: Effect of

Microstructural Modelling,” Mech. Mater., 32, pp. 449–462.

[361] Hanks, B., and Frecker, M., 2019, “Lattice Structure Design for Additive Manufacturing:

Unit Cell Topology Optimization,” ASME International Design Engineering Technical

Conferences & Computers and Information in Engineering, DETC2019-97863.

Page 339: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

310

[362] Fuller, R. B., 1961, “Synergetic Building Construction,” pp. 1–11.

[363] Zheng, X., Lee, H., Weisgraber, T. H., Shusteff, M., DeOtte, J., Duoss, E. B., Kuntz, J. D.,

Biener, M. M., Ge, Q., Jackson, J. A., Kucheyev, S. O., Fang, N. X., and Spadaccini, C.

M., 2014, “Ultralight, Ultrastiff Mechanical Metamaterials,” Science (80-. )., 344(6190),

pp. 1373–1377.

[364] Arabnejad, S., Burnett Johnston, R., Pura, J. A., Singh, B., Tanzer, M., and Pasini, D.,

2016, “High-Strength Porous Biomaterials for Bone Replacement: A Strategy to Assess

the Interplay between Cell Morphology, Mechanical Properties, Bone Ingrowth and

Manufacturing Constraints,” Acta Biomater., 30, pp. 345–356.

[365] Suard, M., 2015, “Characterization and Optimization of Lattice Structures Made by

Electron Beam Melting,” Universite De Grenoble.

[366] de Formanoir, C., Suard, M., Dendievel, R., Martin, G., and Godet, S., 2016, “Improving

the Mechanical Efficiency of Electron Beam Melted Titanium Lattice Structures by

Chemical Etching,” Addit. Manuf., 11, pp. 71–76.

[367] Gangireddy, S., Komarasamy, M., Faierson, E. J., and Mishra, R. S., 2019, “High Strain

Rate Mechanical Behavior of Ti-6Al-4V Octet Lattice Structures Additively

Manufactured by Selective Laser Melting (SLM),” Mater. Sci. Eng. A, 745, pp. 231–239.

[368] Stankovic, T., Mueller, J., and Shea, K., 2017, “The Effect of Anisotropy on the

Optimization of Additively Manufactured Lattice Structures,” Addit. Manuf., 17, pp. 67–

76.

[369] He, Z. Z., Wang, F. C., Zhu, Y. B., Wu, H. A., and Park, H. S., 2017, “Mechanical

Properties of Copper Octet-Truss Nanolattices,” J. Mech. Phys. Solids, 101, pp. 133–149.

[370] International Organization for Standardization, 2011, Mechanical Testing of Metals -

Ductility Testing - Compression Test for Porous and Cellular Metals (ISO 13314:2011.

[371] EOS, “Materials for Metal Additive Manufacturing” [Online]. Available:

https://www.eos.info/material-m. [Accessed: 29-Aug-2019].

[372] Thermo Fisher Scientific, “Avizo” [Online]. Available:

https://www.thermofisher.com/us/en/home/industrial/electron-microscopy/electron-

microscopy-instruments-workflow-solutions/3d-visualization-analysis-software.html.

[373] Hanks, B., Berthel, J., Frecker, M., and Simpson, T., 2020, “Mechanical Properties of

Additively Manufactured Metal Lattice Struturs: Data Review and Design Interface,”

Addit. Manuf., (Accepted for publication).

[374] Dessault Systems, “SolidWorks” [Online]. Available: https://www.solidworks.com/.

[375] Autodesk, “Inventor” [Online]. Available:

https://www.autodesk.com/products/inventor/overview.

[376] PTC, “Creo” [Online]. Available: https://www.ptc.com/en/products/cad/creo.

Page 340: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

311

[377] Materialise, “3-Matic” [Online]. Available: https://www.materialise.com/en/software/3-

matic.

[378] Ashby, M. F., 2017, Materials Selection in Mechanical Design, Butterworth-Heinemann,

Oxford, United Kingdom.

[379] Renishaw, “Data Sheets - Additive Manufacturing” [Online]. Available:

https://www.renishaw.com/en/data-sheets-additive-manufacturing--17862. [Accessed: 29-

Aug-2019].

[380] Gümrük, R., Mines, R. A. W., and Karadeniz, S., 2013, “Static Mechanical Behaviours of

Stainless Steel Micro-Lattice Structures under Different Loading Conditions,” Mater. Sci.

Eng. A, 586, pp. 392–406.

[381] Hedayati, R., Sadighi, M., Mohammadi-Aghdam, M., and Hosseini-Toudeshky, H., 2018,

“Comparison of Elastic Properties of Open-Cell Metallic Biomaterials with Different Unit

Cell Types,” J. Biomed. Mater. Res. - Part B Appl. Biomater., 106(1), pp. 386–398.

[382] Li, S. J., Murr, L. E., Cheng, X. Y., Zhang, Z. B., Hao, Y. L., Yang, R., Medina, F., and

Wicker, R. B., 2012, “Compression Fatigue Behavior of Ti-6Al-4V Mesh Arrays

Fabricated by Electron Beam Melting,” Acta Mater., 60(3), pp. 793–802.

[383] Yavari, S. A., Wauthle, R., Van Der Stok, J., Riemslag, A. C., Janssen, M., Mulier, M.,

Kruth, J. P., Schrooten, J., Weinans, H., and Zadpoor, A. A., 2013, “Fatigue Behavior of

Porous Biomaterials Manufactured Using Selective Laser Melting,” Mater. Sci. Eng. C,

33(8), pp. 4849–4858.

[384] Yavari, S. A., Ahmadi, S. M., Wauthle, R., Pouran, B., Schrooten, J., Weinans, H., and

Zadpoor, A. A., 2015, “Relationship between Unit Cell Type and Porosity and the Fatigue

Behavior of Selective Laser Melted Meta-Biomaterials,” J. Mech. Behav. Biomed. Mater.,

43(March), pp. 91–100.

[385] McKown, S., Shen, Y., Brookes, W. K., Sutcliffe, C. J., Cantwell, W. J., Langdon, G. S.,

Nurick, G. N., and Theobald, M. D., 2008, “The Quasi-Static and Blast Loading Response

of Lattice Structures,” Int. J. Impact Eng., 35(8), pp. 795–810.

[386] Niu, J., Choo, H. L., Sun, W., and Mok, S. H., 2018, “Analytical Solution and

Experimental Study Young’s Modulus of SLM-Fabricated Lattice Structure with

Triangular Unit Cells,” J. Manuf. Sci. Eng.

[387] Al-Saedi, D. S. J., Masood, S. H., Faizan-Ur-Rab, M., Alomarah, A., and Ponnusamy, P.,

2018, “Mechanical Properties and Energy Absorption Capability of Functionally Graded

F2BCC Lattice Fabricated by SLM,” Mater. Des., 144, pp. 32–44.

[388] Choy, S. Y., Sun, C.-N., Leong, K. F., and Wei, J., 2017, “Compressive Properties of Ti-

6Al-4V Lattice Structures Fabricated by Selective Laser Melting: Design, Orientation and

Density,” Addit. Manuf., 16, pp. 213–224.

[389] Ataee, A., Li, Y., Brandt, M., and Wen, C., 2018, “Ultrahigh-Strength Titanium Gyroid

Scaffolds Manufactured by Selective Laser Melting (SLM) for Bone Implant

Page 341: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

312

Applications,” Acta Mater., 158, pp. 354–368.

[390] Beyer, C., and Figueroa, D., 2016, “Design and Analysis of Lattice Structures for Additive

Manufacturing,” J. Manuf. Sci. Eng., 138(12), p. 121014.

[391] Bobbert, F. S. L., Lietaert, K., Eftekhari, A. A., Pouran, B., Ahmadi, S. M., Weinans, H.,

and Zadpoor, A. A., 2017, “Additively Manufactured Metallic Porous Biomaterials Based

on Minimal Surfaces: A Unique Combination of Topological, Mechanical, and Mass

Transport Properties,” Acta Biomater., 53, pp. 572–584.

[392] Borleffs, M. S., 2012, “Finite Element Modeling To Predict Bulk Mechanical Properties

of 3D Printed Metal Foams,” TUDelft.

[393] Campanelli, S. L., Contuzzi, N., Ludovico, A. D., Caiazzo, F., Cardaropoli, F., and Sergi,

V., 2014, “Manufacturing and Characterization of Ti6Al4V Lattice Components

Manufactured by Selective Laser Melting,” Materials (Basel)., 7(6), pp. 4803–4822.

[394] Gümrük, R., and Mines, R. A. W., 2013, “Compressive Behaviour of Stainless Steel

Micro-Lattice Structures,” Int. J. Mech. Sci., 68, pp. 125–139.

[395] Harrysson, O. L. A., Cansizoglu, O., Marcellin-Little, D. J., Cormier, D. R., and West, H.

A., 2008, “Direct Metal Fabrication of Titanium Implants with Tailored Materials and

Mechanical Properties Using Electron Beam Melting Technology,” Mater. Sci. Eng. C,

28(3), pp. 366–373.

[396] Hasib, H. Bin, 2011, “Mechanical Behavior of Non-Stochastic Ti-6Al-4V Cellular

Structures Produced via Electron Beam Melting (EBM),” North Carolina State University.

[397] Heinl, P., Müller, L., Körner, C., Singer, R. F., and Müller, F. A., 2008, “Cellular Ti-6Al-

4V Structures with Interconnected Macro Porosity for Bone Implants Fabricated by

Selective Electron Beam Melting,” Acta Biomater., 4(5), pp. 1536–1544.

[398] Heinl, P., Körner, C., and Singer, R. F., 2008, “Selective Electron Beam Melting of

Cellular Titanium: Mechanical Properties,” Adv. Eng. Mater., 10(9), pp. 882–888.

[399] Leary, M., Maxur, M., Elambasseril, J., Matthew, M., Chirent, T., Sun, Y., Qian, M.,

Easton, M., and Brandt, M., 2016, “Selective Laser Melting (SLM) of AlSi12Mg Lattice

Structures,” Mater. Des., 98, pp. 119–161.

[400] Leary, M., Mazur, M., Williams, H., Yang, E., Alghamdi, A., Lozanovski, B., Zhang, X.,

Shidid, D., Farahbod-Sternahl, L., Witt, G., Kelbassa, I., Choong, P., Qian, M., and

Brandt, M., 2018, “Inconel 625 Lattice Structures Manufactured by Selective Laser

Melting (SLM): Mechanical Properties, Deformation and Failure Modes,” Mater. Des.,

157, pp. 179–199.

[401] Li, C., Lei, H., Liu, Y., Zhang, X., Xiong, J., Zhou, H., and Fang, D., 2018, “Crushing

Behavior of Multi-Layer Metal Lattice Panel Fabricated by Selective Laser Melting,” Int.

J. Mech. Sci., 145(June), pp. 389–399.

[402] Li, S. J., Xu, Q. S., Wang, Z., Hou, W. T., Hao, Y. L., Yang, R., and Murr, L. E., 2014,

Page 342: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

313

“Influence of Cell Shape on Mechanical Properties of Ti-6Al-4V Meshes Fabricated by

Electron Beam Melting Method,” Acta Biomater., 10(10), pp. 4537–4547.

[403] Liu, Y. J., Wang, H. L., Li, S. J., Wang, S. G., Wang, W. J., Hou, W. T., Hao, Y. L.,

Yang, R., and Zhang, L. C., 2017, “Compressive and Fatigue Behavior of Beta-Type

Titanium Porous Structures Fabricated by Electron Beam Melting,” Acta Mater., 126, pp.

58–66.

[404] Liu, Y. J., Li, S. J., Wang, H. L., Hou, W. T., Hao, Y. L., Yang, R., Sercombe, T. B., and

Zhang, L. C., 2016, “Microstructure, Defects and Mechanical Behavior of Beta-Type

Titanium Porous Structures Manufactured by Electron Beam Melting and Selective Laser

Melting,” Acta Mater., 113, pp. 56–67.

[405] Mazur, M., Leary, M., Sun, S., Vcelka, M., Shidid, D., and Brandt, M., 2016,

“Deformation and Failure Behaviour of Ti-6Al-4V Lattice Structures Manufactured by

Selective Laser Melting (SLM),” Int. J. Adv. Manuf. Technol., 84(5–8), pp. 1391–1411.

[406] Mines, R. A. W., Tsopanos, S., Shen, Y., Hasan, R., and McKown, S. T., 2013, “Drop

Weight Impact Behaviour of Sandwich Panels with Metallic Micro Lattice Cores,” Int. J.

Impact Eng., 60, pp. 120–132.

[407] Mullen, L., Stamp, R. C., Brooks, W. K., Jones, E., and Sutcliffe, C. J., 2009, “Selective

Laser Melting: A Regular Unit Cell Approach for the Manufacture of Porous, Titanium,

Bone in-Growth Constructs, Suitable for Orthopedic Applications,” J. Biomed. Mater.

Res. - Part B Appl. Biomater., 89(2), pp. 325–334.

[408] Murr, L. E., Amato, K. N., Li, S. J., Tian, Y. X., Cheng, X. Y., Gaytan, S. M., Martinez,

E., Shindo, P. W., Medina, F., and Wicker, R. B., 2011, “Microstructure and Mechanical

Properties of Open-Cellular Biomaterials Prototypes for Total Knee Replacement Implants

Fabricated by Electron Beam Melting,” J. Mech. Behav. Biomed. Mater., 4(7), pp. 1396–

1411.

[409] Murr, L. E., Gaytan, S. M., Medina, F., Lopez, H., Martinez, E., Machado, B. I.,

Hernandez, D. H., Martinez, L., Lopez, M. I., Wicker, R. B., and Bracke, J., 2010, “Next-

Generation Biomedical Implants Using Additive Manufacturing of Complex, Cellular and

Functional Mesh Arrays,” Philos. Trans. R. Soc. A Math. Phys. Eng. Sci., 368(1917), pp.

1999–2032.

[410] Murr, L., Li, S., Tian, Y., Amato, K., Martinez, E., and Medina, F., 2010, “Open-Cellular

Co-Base and Ni-Base Superalloys Fabricated by Electron Beam Melting,” Materials

(Basel)., 4(4), pp. 782–790.

[411] Parthasarathy, J., Starly, B., Raman, S., and Christensen, A., 2010, “Mechanical

Evaluation of Porous Titanium (Ti6Al4V) Structures with Electron Beam Melting

(EBM),” J. Mech. Behav. Biomed. Mater., 3(3), pp. 249–259.

[412] Ptochos, E., and Labeas, G., 2012, “Elastic Modulus and Poisson ’ s Ratio Determination

of Micro-Lattice Cellular Structures by Analytical , Numerical and Homogenisation

Methods,” J. Sandw. Struct. Mater., 14(5).

Page 343: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

314

[413] Rehme, O., and Emmelmann, C., 2006, “Rapid Manufacturing of Lattice Structures with

Selective Laser Melting,” Laser-based Micropackaging, 6107(February 2006), p. 61070K.

[414] Sallica-Leva, E., Jardini, A. L., and Fogagnolo, J. B., 2013, “Microstructure and

Mechanical Behavior of Porous Ti-6Al-4V Parts Obtained by Selective Laser Melting,” J.

Mech. Behav. Biomed. Mater., 26, pp. 98–108.

[415] Shen, Y., McKown, S., Tsopanos, S., Sutcliffe, C. J., Mines, R. A. W., and Cantwell, W.

J., 2010, “The Mechanical Properties of Sandwich Structures Based on Metal Lattice

Architectures,” J. Sandw. Struct. Mater., 12(2), pp. 159–180.

[416] Tsopanos, S., Mines, R. A. W., McKown, S., Shen, Y., Cantwell, W. J., Brooks, W., and

Sutcliffe, C. J., 2010, “The Influence of Processing Parameters on the Mechanical

Properties of Selectively Laser Melted Stainless Steel Microlattice Structures,” J. Manuf.

Sci. Eng., 132, pp. 1–12.

[417] van Grunsven, W., Hernandez-Nava, E., Reilly, G., and Goodall, R., 2014, “Fabrication

and Mechanical Characterisation of Titanium Lattices with Graded Porosity,” Metals

(Basel)., 4(3), pp. 401–409.

[418] Wauthle, R., Vrancken, B., Beynaerts, B., Jorissen, K., Schrooten, J., Kruth, J.-P., and

Van Humbeeck, J., 2015, “Effects of Build Orientation and Heat Treatment on the

Microstructure and Mechanical Properties of Selective Laser Melted Ti6Al4V Lattice

Structures,” Addit. Manuf., 5, pp. 77–84.

[419] Xiao, L., Song, W., Wang, C., Liu, H., Tang, H., and Wang, J., 2015, “Mechanical

Behavior of Open-Cell Rhombic Dodecahedron Ti–6Al–4V Lattice Structure,” Mater. Sci.

Eng. A, 640, pp. 375–384.

[420] Yan, C., Hao, L., Hussein, A., and Young, P., 2015, “Ti-6Al-4V Triply Periodic Minimal

Surface Structures for Bone Implants Fabricated via Selective Laser Melting,” J. Mech.

Behav. Biomed. Mater., 51, pp. 61–73.

[421] Yánez, A., Cuadrado, A., Martel, O., Afonso, H., and Monopoli, D., 2018, “Gyroid Porous

Titanium Structures: A Versatile Solution to Be Used as Scaffolds in Bone Defect

Reconstruction,” Mater. Des., 140, pp. 21–29.

[422] Yang, E., Leary, M., Lozanovski, B., Downing, D., Mazur, M., Sarker, A., Khorasani, A.,

Jones, A., Maconachie, T., Bateman, S., Easton, M., Qian, M., Choong, P., and Brandt,

M., 2019, “Effect of Geometry on the Mechanical Properties of Ti-6Al-4V Gyroid

Structures Fabricated via SLM: A Numerical Study,” Mater. Des., 184, p. 108165.

[423] Zaharin, H. A., Rani, A. M. A., Azam, F. I., Ginta, T. L., Sallih, N., Ahmad, A., Yunus, N.

A., and Zulkifli, T. Z. A., 2018, “Effect of Unit Cell Type and Pore Size on Porosity and

Mechanical Behavior of Additively Manufactured Ti6Al4V Scaffolds,” Materials (Basel).,

11(12).

[424] Zhao, S., Li, S. J., Hou, W. T., Hao, Y. L., Yang, R., and Misra, R. D. K., 2016, “The

Influence of Cell Morphology on the Compressive Fatigue Behavior of Ti-6Al-4V Meshes

Fabricated by Electron Beam Melting,” J. Mech. Behav. Biomed. Mater., 59, pp. 251–264.

Page 344: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

315

[425] Zhao, M., Liu, F., Fu, G., Zhang, D. Z., Zhang, T., and Zhou, H., 2018, “Improved

Mechanical Properties and Energy Absorption of BCC Lattice Structures with Triply

Periodic Minimal Surfaces Fabricated by SLM,” Materials (Basel)., 11(12).

[426] Zhu, H. X., Knott, J. F., and Mills, N. J., 1997, “Analysis of the Elastic Properties of

Open-Cell Foams with Tetrakaidecahedral Cells,” J. Mech. Phys. Solids, 45(3), pp. 319–

343.

[427] Rohatgi, A., 2019, “WebPlotDigitizer” [Online]. Available:

https://automeris.io/WebPlotDigitizer.

[428] Drevon, D., Fursa, S. R., and Malcolm, A. L., 2017, “Intercoder Reliability and Validity

of WebPlotDigitizer in Extracting Graphed Data,” Behav. Modif., 41(2), pp. 323–339.

[429] Moeyaert, M., Maggin, D., and Verkuilen, J., 2016, “Reliability, Validity, and Usability of

Data Extraction Programs for Single-Case Research Designs,” Behav. Modif., 40(6), pp.

874–900.

[430] Burda, B. U., O’Connor, E. A., Webber, E. M., Redmond, N., and Perdue, L. A., 2017,

“Estimating Data from Figures with a Web-Based Program: Considerations for a

Systematic Review,” Res. Synth. Methods, 8(3), pp. 258–262.

[431] McBride, G. B., 2005, “A Proposal for Strength-of-Agreement Criteria for Lins

Concordance Correlation Coefficient,” NIWA Client Rep., HAM2005-06.

[432] 2019, “Desmos Online Calculator” [Online]. Available:

https://www.desmos.com/calculator/7yuojxowmx. [Accessed: 10-Jan-2019].

[433] Maxwell, J. C., 1864, “XLV. On Reciprocal Figures and Diagrams of Forces,” London,

Edinburgh, Dublin Philos. Mag. J. Sci., 27(182), pp. 250–261.

Page 345: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

316

Appendix A: MATLAB Image Analysis

For image analysis in MATLAB, this file is set up to read in multiple images and

calculate the f_over and f_under objective functions from thermal ablation model. Special thanks

to Zackary Snow for his support in recommending and providing this direction of image analysis

for computing the objective functions.

% This file computes the f_over and f_under objective function for an

% ablation zone. The image analysis is performed on three different cross

% sections of the ablation that are output from COMSOL.

%Wipes previously stored variables, cleans workspace, and closes open figures

clear; %clc;

files = char('image2_45_600.tiff', 'image2_xz_600.tiff', 'image2_yz_600.tiff');

pixels_to_mm = 1/36;

abl_rad = 12.5; % Target ablation radius (mm)

angle_res = .15; % Take measurements every 'angle_res' degrees

angle_tol = .15; % Take measurements at each angle + or - the tolerance

%% DATA ANALYSIS %%

%For loop to analyze each image file

nfiles = size(files,1);

fover = zeros(nfiles,2);

funder = zeros(nfiles,2);

for k = 1:nfiles

% Imports an image file and converts to grayscale and inverts colors

filename = files(k,:);

[A,map] = imread(filename);

grayscale_version = imcomplement(rgb2gray(A));

% Converts grayscale image to black and white image

thresholded_image = imbinarize(grayscale_version);

% Analyzes the properties of the ablation zone in the image

image_stats = regionprops(thresholded_image,'BoundingBox');

bounding_box = cat(1,image_stats.BoundingBox);

%% UNGULATION COUNTING AND QUANTIFICATION ALGORITHM %%

% Establishes the bounding box around the ablation zone

x_lower = int16(round(bounding_box(1)));

x_upper = int16(fix(bounding_box(1)))+int16(round(bounding_box(3)));

x_mid = double((x_upper+x_lower)/2);

x_range = x_lower:x_upper;

y_lower = int16(round(bounding_box(2)));

y_upper = int16(fix(bounding_box(2)))+int16(round(bounding_box(4)));

y_mid = double((y_upper+y_lower)/2);

Page 346: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

317

y_range = y_lower:y_upper;

% Generates smaller image of the bounding box around each ablation zone

bb_BW = thresholded_image( y_range(1)-1:y_range(end), x_range(1)-

1:x_range(end)+1);

% Calculates the distance from and angle relative to the center of

% area of the particle for each pixel contained within the boundning

% box of the particle.

distance = zeros(length(y_range),length(x_range));

phi = zeros(length(y_range),length(x_range));

for j = 1:length(y_range)

for i = 1:length(x_range)

distance(j,i) = ((double(i-1+x_lower-x_mid))^2 + (double(j-

1+y_lower-y_mid))^2)^(0.5);

phi(j,i) = -1*atan2d(double(j-1+y_lower-y_mid),double(i-

1+x_lower-x_mid));

if phi(j,i)<0

phi(j,i) = phi(j,i)+360;

else

phi(j,i) = phi(j,i);

end

if bb_BW(j,i) == 0 % If pixel is black, set distance to zero

distance(j,i) = 0;

else

distance(j,i) = distance(j,i);

end

end

end

%'edge_distance' is the distance from the edge of the particle from

%the center of area of the particle.

edge_distance = zeros(1,round(360/angle_res));

count = 0;

for angle = 0:angle_res:(360-angle_res)

count = count + 1;

positions = find(phi > (angle - angle_tol) & phi < (angle +

angle_tol));

if isempty(positions)

edge_distance(1,count) = 0;

else

edge_distance(1,count) = max(distance(positions))*pixels_to_mm;

end

end

% Computes the differences between the edge distance (from center

% to edge) and the target ablation radius

Distance_Difference = edge_distance - abl_rad;

ind_over = Distance_Difference > 0;

ind_under = Distance_Difference < 0;

fover(k,:) = [sum(Distance_Difference(ind_over))/abl_rad,

length(Distance_Difference(ind_over))];

funder(k,:) = [abs(sum(Distance_Difference(ind_under))/abl_rad),

length(Distance_Difference(ind_under))];

end

Page 347: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

318

f_o = sum(fover(:,1))/sum(fover(:,2))

f_u = sum(funder(:,1))/sum(funder(:,2))

if isnan(f_o)

f_o = 0;

end

if isnan(f_u)

f_u = 0;

end

Page 348: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

319

Appendix B: ALTO 3D MATLAB Files

Main File

The main file is used to set up the optimization problem and contains essentially all the

parameters to determine what type of optimization will be run. It calls the execute function to

execute the optimization problem.

ALTO3D_main.m

% Multi-objective unit cell generation

% Written by Brad Hanks

%

% This file contains all of the preliminary design and input information

% required from the user for the development of the optimization.

clear; close all; clc; % Begin with a fresh slate

% Specify the size of the ground structure for optimization

opt.ucs = [60,20,20]; % Unit cell size in mm

opt.gssize = [ 7, 2, 5]; % Number of nodes in [ x, y, z]

% Specify the objective functions and constraints

% opt.objs = [ TYPE, MINIMIZE, WEIGHT, TARGET]

% TYPE: 1 - calculated the strain energy

% 2 - calculates the thermal conductivity between a set of nodes

% 3 - calculates the homogenized unit cell properties

% 4 - calculates the total volume of the material

% 5 - Number of relevant members

% 6 - Powder Removability (Maxi Avg Dist, Maxi Min Dist, Mini #

trapped regions

% MINIMIZE: 1 to minimize the objective

% -1 to maximize the objective

% Should be 1 for objectives 3 and 4

% WEIGHT: specify weight value or set of weight values

% leave blank for general pareto sweep

% TARGET: Specify target value for objective

% For powder removability, if maximize:

% 1 = maximize the avg distance of between relevant members

% 2 = maximize the min distance between relevant members

% For powder removability, if minimize, provide tolerance here

% NOTE: For coupled strain energy and conductivity, the conductivity

% objective function must follow immediately after the strain energy

% objective function.

% Homogenization Target Options

% Isotropic

% E = 0.8*2e5;

% nu = 0.3;

% target = E/(1-nu^2)*[1, nu, 0;

Page 349: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

320

% nu, 1, 0;

% 0, 0, (1-nu)/2];

% % Orthotropic

% factor = 3;

% opt.E1 = 0.8*2e5;

% opt.E2 = opt.E1/factor;

% opt.nu12 = 0.6;

% opt.nu21 = opt.nu12/factor;

% opt.G12 = opt.E1/factor/(1+opt.nu12);

% homog_target = [ opt.E1/(1-opt.nu12*opt.nu21), opt.nu21*opt.E1/(1-

opt.nu12*opt.nu21), 0;

% opt.nu12*opt.E2/(1-opt.nu12*opt.nu21), opt.E2/(1-

opt.nu12*opt.nu21), 0;

% 0, 0, opt.G12];

% % Match Diamond UCS 8mm, Diam 1mm

% val = (1/2)*1.8944e3;

% homog_target = zeros(6,6);

% homog_target([1,2,3,7,8,9,13,14,15]) = val;

% Match Octet Truss UCS 8mm, Diam 1mm

val = 6.9420e3;

homog_target = zeros(6,6);

homog_target([1,8,15]) = val;

homog_target([2,3,7,9,13,14]) = val/2;

homog_target([22,29,36]) = 0.625*val;

opt.objs(1) = struct('type', 1,'minimize',

1,'scale',[1],'weight',[1.0],'target',[]);

% opt.objs(1) = struct('type', 2,'minimize',-

1,'scale',[1],'weight',[1.0],'target',[0]);

% opt.objs(1) = struct('type',

3,'minimize',1,'scale',[1],'weight',[1.0],'target',homog_target);

% opt.objs(2) = struct('type',

4,'minimize',1,'scale',[1],'weight',[1.0],'target',[0.01]); % The target is a

volume fraction and will OVERRIDE opt.volfrac

% opt.objs(3) = struct('type',

5,'minimize',1,'scale',[1],'weight',[1.0],'target',[]);

% opt.objs(4) = struct('type',

6,'minimize',1,'scale',[1],'weight',[1.0],'target',4);

% Specify Optimization constraint parameters

opt.volfrac = [.58]; % Volume Fraction constraint, leave empty to remove

constraints

opt.dlim = [9.5e-5, 7.5]; % Element diameter limits [min,max] in mm

opt.setsym = ['y']; % Set symmetry mid planes { X, Y, Z, XY, XZ, XYZ,

NONE }

opt.buildang = []; % Build Angle Limit measured from horizontal

opt.removenode = []; % Remove elements connected to these nodes if row

vector, column vector removes only the element between the two nodes;

% ~~ 4x4x4 nodes for common structures ~~

% ~~ 5x5x5 nodes for common structures ~~

% Diamond List -

[2,4,6:10,12,14,16:20,22,24,26:31,35,36,40,41,45:50,52,54,56,60,66,70,72,74,76:

81,85,86,90,91,95,96:100,102,104,106:110,112,114,116:120,122,124,1,3,11,15,23,2

5,51,55,71,75,103,105,111,115,121,123]

% Diamond List2-

[2,4,6:10,12,14,16:20,22,24,26:31,35,36,40,41,45:50,52,54,56,60,66,70,72,74,76:

Page 350: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

321

81,85,86,90,91,95,96:100,102,104,106:110,112,114,116:120,122,124,3,5,11,15,21,2

3,51,55,71,75,101,103,111,115,123,125]

% Octet List -

[2,3,4,6:10,11,12,14,15,16:20,22,23,24,26:31,35,36,40,41,45:50,51,52,54,55,56,6

0,66,70,71,72,74,75,76:81,85,86,90,91,95,96:100,102,103,104,106:110,111,112,114

,115,116:120,122,123,124]

% 9pts per face -

[2,4,6:10,12,14,16:20,22,24,26:31,35,36,40,41,45:50,52,54,56,60,66,70,72,74,76:

81,85,86,90,91,95,96:100,102,104,106:110,112,114,116:120,122,124]

% Interior List - [32:34,37:39,42:44,57:59,62:64,67:69,82:84,87:89,92:94]

% 2,4 Edge -

[1,3,5,7:9,11,12,14,15,17:19,21,23,25,27:29,31,35,36,40,41,45,47,48,49,51,52,54

,55,56,60,66,70,71,72,74,75,77:79,81,85,86,90,91,95,97,98,99,101,103,105,107:10

9,111,112,114,115,117:119,121,123,125]

% Edge Only -

[7:9,12:14,17:19,27:29,31,35,36,40,41,45,47:49,52:54,56,60,61,65,66,70,72:74,77

:79,81,85,86,90,91,95,97:99,107:109,112:114,117:119];

opt.removesurface = false; % Remove surface elements

opt.removelength = []; % Remove members whose length falls between this range

% Penalty functions

opt.drel = [1]; % Relevant diameter (mm) - Used to determine which members

are relevant

opt.Pcount = []; % '+' few members, '-' more members

opt.Pdmid = []; % '+' few members between drel and dlim(1)

opt.Ptype = 1; %1 - None, 2 - overlap, 3 - connectivity, 4 - both

opt.SPtype = 1; %1 - None, 2 - neighbor, 3 - self Support Structure

Requirements

% Specify material properties

opt.mod = 2e5; % Modulus of elasticity N/mm^2

opt.kcond = 0.054; % Thermal conductivity W/mm-K

opt.alpha = 12e-6; % Coefficient of thermal expansion mm/(mm-K)

% Specify optimization options

opt.algorithm = 'fmincon'; % Specify algorithm type, fmincon or borg or NSGA-

II

switch lower(opt.algorithm)

case {'borg'}

opt.epsilon = [1]; % Specify objective tolerances

opt.FuncCount = 1000; % Specify max function evaluations for borg

case {'fmincon'}

opt.parallel = false; % Run in parallel

opt.MaxIterations = 500; % Maximum number of iterations, enter

1 for single run and not optimization

opt.minstep = 1e-6; % Step size tolerance, default 1e-10

opt.saveobj = 0; % Save all objective function values

(EXPENSIVE)

case {'nsga-ii'}

opt.FunctionTolerance = 1e-4; % Tolerance for relative spread change

for convergence over MaxStall Generations, default 1e-4

opt.MaxIterations = 2449; % Maximum Generations, default is

200*numvars, To define by FuncCount-- ((FuncCount/Population) - 1)

opt.MaxStallGenerations = 100; % Number of stall generations

considered in convergence, default is 100

opt.ParetoFraction = 0.35; % Number of individuals to keep from

first Pareto Front, default is 0.35

opt.PopulationSize = 200; % Population size, default 50 (numvars

<= 5), 200 (numvars > 5)

Page 351: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

322

opt.UseParallel = true; % Use Parallel computing when available

opt.CrossoverFraction = 0.80;

end

% Initial Population options

opt.init = []; % Populate initial condition *ROW vector, will

randomize if blank

% opt.init = pi/4*(opt.dlim(1)^2)*ones(1,2016); % Populate initial

condition at A_lower

% opt.init = pi/4*(opt.drel^2)*ones(1,2016); % Populate initial

condition at A_rel

% opt.init = pi/4*(opt.dlim(2)^2)*ones(1,7750); % Populate initial

condition at A_upper

% opt.init([5,45,66,69,79,100,102,110])=1.0;

% Specify the type of loading and boundary conditions

% Required for strain energy objective function

% opt.load syntax: [ node#, fx, fy, fz;

% node#, fx, fy, fz];

opt.loads = [ 13, 0, 0, 1000;

14, 0, 0, 1000];

% opt.bc syntax: [ node#, ux, uy, uz;

% node#, ux, uy, uz];

% ux = 1 is free, ux = 0 is constrained

opt.bcs = [ 1, 0, 0, 0;

2, 0, 0, 0;

15, 0, 0, 0;

16, 0, 0, 0;

29, 0, 0, 0;

30, 0, 0, 0];

% Specify temperature and nodes for thermal conduction objective function

% Required for thermal conduction objective function

% in = [1:12,14:31,35,36,40,41,45:56,60,61,65,66,70:81,85,86,90,91,95:125]; %

Nodes at temperature T1

% out = [13]; % Nodes at temperature T2

in = [];

out = [53];

Iin = 5*ones(124,1);

opt.temps = [ 100, 25, 25]; % [T1, T2, T-reference] [degC]

opt.T1nodes = in;

opt.T2nodes = out;

opt.Iin = Iin;

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%

%%

%Single Optimization

modopt = []; % Leave empty for non-parallel runs

filename = 'SE_bench_R4_dr_10_5';

ALTO3D_exec(opt,filename,modopt)

% % Parallel Runs

% nprocs = 3;

% parfor(modopt = 1:3,nprocs)

Page 352: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

323

% filename = ['C1_R5_AM_eps2_M',num2str(modopt)]; % Filename prefix, do not

include ".mat"

% ALTO3D_exec(opt,filename,modopt)

% end

Execution Function

The execution file takes all the parameters from the problem set in the main file and the

initializes the ground structure, applies ground structure restrictions, and prepares variables for

the optimization. It then sends the problem to the specified optimization algorithm.

ALTO3D_exec.m

% Multi-objective unit cell generation

% Written by Brad Hanks

%

% This file reads all of the input information from the main file and

% initializes the ground structure accordingly. After initializing the

% ground structure, this file executes the optimization.

function [] = ALTO3D_exec(opt,filename,modopt)

% Setup filename

opt.filename = filename;

% Setup different modifications for parfor loop here

if modopt == 1

disp('Running configuration 1 - (1/2x) no DfAM')

else

disp('Running main file config')

end

% Initialize Ground Structure Geometry

gs = ALTO3D_initgs(opt);

A = gs.Ainitial;

% Initialize

% Determine objective functions

optim.numobjs = length([opt.objs.type]);

for i = 1:optim.numobjs

if opt.objs(i).type == 1

disp('Initializing Strain Energy ObjFunc')

optim.compl.loads = opt.loads;

optim.compl.bcs = opt.bcs;

optim.compl.objnum = i;

% Calculate range of objective function

opt.type = 1;

optim.compl.range(1) =

nonzeros(ALTO3D_compltherm(gs.Aupper,opt,optim,gs));

optim.compl.range(2) =

nonzeros(ALTO3D_compltherm(gs.Alower,opt,optim,gs));

elseif opt.objs(i).type == 2

disp('Initializing Thermal Conductivity ObjFunc')

Page 353: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

324

optim.therm.T1 = opt.temps(1);

optim.therm.T2 = opt.temps(2);

optim.therm.Tref = opt.temps(3);

optim.therm.in = opt.T1nodes;

optim.therm.out = opt.T2nodes;

% Calculate range of objective function

opt.type = 2;

optim.therm.range(1) =

nonzeros(ALTO3D_compltherm(gs.Alower,opt,optim,gs));

optim.therm.range(2) =

nonzeros(ALTO3D_compltherm(gs.Aupper,opt,optim,gs));

elseif opt.objs(i).type == 3

disp('Initializing Homogenization Properties')

optim.homog.target = opt.objs(i).target;

elseif opt.objs(i).type == 4

disp('Initializing Volume ObjFunc')

optim.volume.range(1) = gs.Alower*gs.L;

optim.volume.range(2) = gs.Aupper*gs.L;

optim.volume.vftarget = opt.objs(i).target;

elseif opt.objs(i).type == 5

disp('Initializing Member Count ObjFunc')

elseif opt.objs(i).type == 6

if opt.objs(i).minimize == 1

optim.powder.typename = 'Minimize Trapped Regions';

optim.powder.type = 3;

optim.powder.tol = opt.objs(i).target;

else

if opt.objs(i).target == 1

optim.powder.typename = 'Maximize Avg Distance';

optim.powder.type = 1;

elseif opt.objs(i).target == 2

optim.powder.typename = 'Maximize Min Distance';

optim.powder.type = 2;

else

error('Unknown Powder Removability Target')

end

end

disp(['Initializing Powder Removability ObjFunc:

',[optim.powder.typename]])

optim.powder.target = opt.objs(i).target;

else

error(['Unknown objective function type, TYPE:

',num2str(opt.objs(i).type)])

end

end

% % Check Code

% gs.Aupper = 0.04*ones(1,gs.M);

% opt.type = 12;

% ALTO3D_compltherm(gs.Aupper,opt,optim,gs)

% Objective Function Weights

try

optim.weight = vertcat(opt.objs.weight)';

catch

error('Weights are not the same length for each objective')

end

if isempty(optim.weight)

if optim.numobjs == 2

optim.weight = [ 1.0:-0.1:0.0; 0.0:0.1:1.0]';

Page 354: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

325

elseif optim.numobjs == 3

optim.weight = [ 1, 0, 0, 0.5, 0.5, 0, (1/3), 0.4, 0.4, 0.2;

0, 1, 0, 0.5, 0, 0.5, (1/3), 0.4, 0.2, 0.4;

0, 0, 1, 0, 0.5, 0.5, (1/3), 0.2, 0.4, 0.4]';

else

error('Requires weight values to be defined')

end

end

%% Prepare Optimization

% Optimization variables

optim.Ainitial = gs.Ainitial;

optim.L = gs.L;

optim.IEN = gs.IEN;

optim.Alower = gs.Alower;

optim.Aupper = gs.Aupper;

optim.theta = gs.theta;

optim.thetab = gs.thetab;

if ~isempty(opt.volfrac)

% optim.Vstar = opt.volfrac*(optim.Aupper*gs.L);

optim.Vstar = opt.volfrac*(opt.ucs(1)*opt.ucs(2)*opt.ucs(3));

else

optim.Vstar = [];

end

optim.wght = [];

optim.en_symmap = zeros(gs.M,1);

optim.en_count = zeros(gs.M,1);

optim.symm = [];

optim.keep = [];

optim.remove = [];

%% Reduce number of elements

% Remove elements by symmetry

for i = 1:gs.N

symnodes(1,i) = find(abs(gs.coord(1,:)-(opt.ucs(1)-gs.coord(1,i)))<.00001 &

gs.coord(2,:) == gs.coord(2,i) & gs.coord(3,:) == gs.coord(3,i));

symnodes(2,i) = find(abs(gs.coord(2,:)-(opt.ucs(2)-gs.coord(2,i)))<.00001 &

gs.coord(3,:) == gs.coord(3,i) & gs.coord(1,:) == gs.coord(1,i));

symnodes(3,i) = find(abs(gs.coord(3,:)-(opt.ucs(3)-gs.coord(3,i)))<.00001 &

gs.coord(1,:) == gs.coord(1,i) & gs.coord(2,:) == gs.coord(2,i));

end

IEN1 = zeros(2,gs.M); % Flip across X-mid

IEN2 = zeros(2,gs.M); % Flip across Y-mid

IEN3 = zeros(2,gs.M); % Flip across Z-mid

IEN4 = zeros(2,gs.M); % Flip across X-mid then Y-mid

IEN5 = zeros(2,gs.M); % Flip across Z-mid then X-mid

IEN6 = zeros(2,gs.M); % Flip across Z-mid then Y-mid

IEN7 = zeros(2,gs.M); % Flip across Z-mid then X-mid then Y-mid

for i = 1:gs.N

map1 = gs.IEN == i;

IEN1(map1) = symnodes(1,i);

IEN2(map1) = symnodes(2,i);

IEN3(map1) = symnodes(3,i);

end

for i = 1:gs.N

map2 = IEN1 == i;

Page 355: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

326

IEN4(map2) = symnodes(2,i);

end

for i = 1:gs.N

map3 = IEN3 == i;

IEN5(map3) = symnodes(1,i);

IEN6(map3) = symnodes(2,i);

end

for i = 1:gs.N

map4 = IEN5 == i;

IEN7(map4) = symnodes(2,i);

end

IEN1 = sort(IEN1);

IEN2 = sort(IEN2);

IEN3 = sort(IEN3);

IEN4 = sort(IEN4);

IEN5 = sort(IEN5);

IEN6 = sort(IEN6);

IEN7 = sort(IEN7);

switch lower(opt.setsym)

case{'x'}

IEN_f = [gs.IEN;IEN1];

case{'y'}

IEN_f = [gs.IEN;IEN2];

case{'z'}

IEN_f = [gs.IEN;IEN3];

case{'xy'}

IEN_f = [gs.IEN;IEN1;IEN2;IEN4];

case{'xz'}

IEN_f = [gs.IEN;IEN1;IEN3;IEN5];

case{'yz'}

IEN_f = [gs.IEN;IEN2;IEN3;IEN6];

case{'xyz'}

IEN_f = [gs.IEN;IEN1;IEN2;IEN3;IEN4;IEN5;IEN6;IEN7];

case{'none'}

IEN_f = [gs.IEN];

end

gs.IEN_sym = zeros(2,gs.M);

for i = 1:gs.M

row = find(IEN_f(:,i)==min(IEN_f(1:2:(size(IEN_f,1)-1),i)));

if length(row)>1

min_pos = find(IEN_f(row+1,i)==min(IEN_f(row+1,i)));

row = row(min_pos(1));

end

gs.IEN_sym(:,i) = IEN_f([row,row+1],i);

end

elem = 1;

for i = 1:gs.M

if optim.en_symmap(i) <1

nodeA = gs.IEN_sym(1,i);

nodeB = gs.IEN_sym(2,i);

e_pos = find(gs.IEN_sym(1,:) == nodeA & gs.IEN_sym(2,:) == nodeB);

optim.en_symmap(e_pos) = elem;

optim.en_count(e_pos) = length(e_pos); %Add count of number of elements

repeated

Page 356: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

327

elem = elem + 1;

if length(e_pos) > 1

optim.symm = [optim.symm,e_pos(2:end)];

end

end

end

optim.Ainitial(optim.symm) = [];

optim.en_count(optim.symm) = [];

optim.L(optim.symm) = [];

optim.Alower(optim.symm) = [];

optim.Aupper(optim.symm) = [];

optim.theta(optim.symm,:) = [];

optim.thetab(optim.symm) = [];

optim.IEN(:,optim.symm) = [];

% Remove elements from optimization that will be set to lower limit

for i = 1:length(optim.L)

% Remove based on build angle

if ~isempty(opt.buildang)

if ((abs(optim.thetab(i)) > pi*opt.buildang(1)/180) &&

(abs(optim.thetab(i)) < pi*opt.buildang(2)/180))

optim.remove = [optim.remove;i];

continue;

end

end

% Remove based on nodal connections

if ~isempty(opt.removenode)

% if any(optim.IEN(:,i) == opt.removenode, 'all') % MIN VERSION 2018b

if any(any(optim.IEN(:,i) == opt.removenode))

optim.remove = [optim.remove;i];

continue;

end

end

% Remove surface elements

if opt.removesurface

coordA = gs.coord(:,optim.IEN(1,i));

coordB = gs.coord(:,optim.IEN(2,i));

surface = [coordA,coordB]==[min(gs.coord,[],2),min(gs.coord,[],2)];

surface2 = [coordA,coordB]==[max(gs.coord,[],2),max(gs.coord,[],2)];

if any(and(surface(:,1),surface(:,2))) ||

any(and(surface2(:,1),surface2(:,2)))

optim.remove = [optim.remove;i];

continue;

end

end

% Remove based on element length

if ~isempty(opt.removelength)

if ((optim.L(i) > opt.removelength(1)) && (optim.L(i) <

opt.removelength(2)))

optim.remove = [optim.remove;i];

continue;

end

end

optim.keep = [optim.keep;i];

end

Page 357: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

328

optim.Ainitial(optim.remove) = [];

optim.en_count(optim.remove) = [];

optim.L(optim.remove) = [];

optim.Alower(optim.remove) = [];

optim.Aupper(optim.remove) = [];

optim.theta(optim.remove,:) = [];

optim.thetab(optim.remove) = [];

optim.IEN(:,optim.remove) = [];

% %% Visualize Prior to Optimization

% Ain = optim.Ainitial;

% Areduced_ck = optim.Ainitial;

% Areduced_ck([optim.keep]) = Areduced_ck;

% Areduced_ck([optim.remove]) = gs.Alower(1);

%

% A_ck = zeros(1,gs.M);

% if isfield(optim,'en_symmap')

% for i = 1:gs.M

% A_ck(i) = Areduced_ck(optim.en_symmap(i));

% if ~isempty(opt.removenode)

% if any(any(gs.IEN(:,i) == opt.removenode))

% A_ck(i) = gs.Alower(1);

% end

% end

% end

% end

% ALTO3D_plot(A_ck,gs.IEN,gs.coord,.5)

% return

%%% Execute Optimization %%%

switch lower(opt.algorithm)

case{'fmincon'}

Results = zeros(size(optim.weight,1),size(optim.weight,2)+1);

for run = 1:size(optim.weight,1)

optim.wght = optim.weight(run,:);

%% Save current

if ~isempty(opt.filename)

opt.tempname = [opt.filename,'-wghts',sprintf('_%.2f',

optim.wght),'.mat'];

if exist(opt.tempname,'file')

error('Change filename to avoid overwiting')

end

outvars =

struct('Iteration',[1],'FuncCount',[],'A',[],'vol',[],'fval',[],'objs',[],'P',[

]);

save(opt.tempname,'opt','gs','optim','outvars')

end

f = @(A)ALTO3D_ucsim(A,opt,optim,gs);

optout = @(x,optimValues,state)

ALTO3D_out(x,optimValues,state,opt); % Pass vars to output fcn

options =

optimoptions(@fmincon,'StepTolerance',opt.minstep,'MaxIterations',opt.MaxIterat

ions,'PlotFcns',@optimplotfval,'MaxFunctionEvaluations',2*opt.MaxIterations*len

gth(optim.Ainitial),'OutputFcn',

{optout},'UseParallel',opt.parallel,'OptimalityTolerance',1e-6);

Page 358: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

329

fprintf('Initializing Optimization at %s \n',datestr(now)); tic;

fprintf('Time \t\t Iter \t Fval \n')

[A,fval,exitflag,output] =

fmincon(f,optim.Aupper,((optim.L).*optim.en_count)',optim.Vstar,[],[],optim.Alo

wer,optim.Aupper,[],options);

% [A,fval,exitflag,output] =

fmincon(f,optim.Ainitial,[],[],[],[],optim.Alower,optim.Aupper,[],options);

fprintf('Optimization Completed at %s \n',datestr(now))

Results(run,:) = [optim.wght, fval];

time = toc;

save(opt.tempname,'time','-append')

end

case{'borg'}

optim.wght = ones(1,optim.numobjs);

if ~isempty(opt.filename)

if exist([opt.filename,'.mat'],'file')

warning(['Tried to overwrite ',opt.filename,'.mat'])

warning('Change Filename to avoid overwrite')

return

elseif exist([opt.filename],'file')

warning('Change Filename to avoid overwrite')

disp(['Tried to overwrite ',opt.filename,'.mat'])

return

else

save(opt.filename,'opt','gs','optim')

end

end

f = @(A)ALTO3D_ucsim(A,opt,optim,gs);

disp(['ModOpt ',num2str(modopt),' Vars -

',num2str(length(optim.Ainitial))])

fprintf('Initialilzing BORG Optimization at %s \n',datestr(now)); tic;

[vars, objs] =

borg(length(optim.Ainitial),optim.numobjs,1,f,opt.FuncCount,opt.epsilon,optim.A

lower,optim.Aupper)

fprintf('Optimization Terminated at %s \n',datestr(now))

time = toc;

save(opt.filename,'vars','objs','time','-append')

case{'nsga-ii'}

optim.wght = ones(1,optim.numobjs);

if ~isempty(opt.filename)

if exist([opt.filename,'.mat'],'file')

error('Change Filename to avoid overwrite')

elseif exist([opt.filename],'file')

error('Change Filename to avoid overwrite')

else

save(opt.filename,'opt','gs','optim')

end

end

f = @(A)ALTO3D_ucsim(A,opt,optim,gs);

optout = @(options,state,flag)ALTO3D_gaout(options,state,flag,opt); %

Pass vars to output fcn

options =

optimoptions(@gamultiobj,'FunctionTolerance',opt.FunctionTolerance,'MaxGenerati

Page 359: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

330

ons',opt.MaxIterations,'MaxStallGenerations',opt.MaxStallGenerations,'ParetoFra

ction',opt.ParetoFraction,'PopulationSize',opt.PopulationSize,'UseParallel',opt

.UseParallel,'OutputFcn',{optout},'CrossoverFraction',opt.CrossoverFraction,'Us

eParallel',opt.UseParallel);

fprintf('Intializing NSGA-II Optimization at %s \n',datestr(now)); tic;

if isempty(optim.Vstar)

[x, fval, exitflag, output, population, scores] =

gamultiobj(f,length(optim.Ainitial),[],[],[],[],optim.Alower,optim.Aupper,optio

ns);

else

[x, fval, exitflag, output, population, scores] =

gamultiobj(f,length(optim.Ainitial),((optim.L).*optim.en_count)',optim.Vstar,[]

,[],optim.Alower,optim.Aupper,options);

end

fprintf('Optimization Completed at %s \n',datestr(now));

time=toc;

save(opt.filename,'x','fval','exitflag','output','population','scores','time','

-append')

end

end

Ground Structure Initialization Function

The ground structure initialization function initializes the entire ground structure

according to the optimization problem specifications. It also sets up the node-neighbor pairs for

neighboring unit cells and reads in the overlapping elements matrix.

ALTO3D_initgs.m

% Multi-objective unit cell generation

% Written by Brad Hanks

%

% This file reads all of the input information from the main file and

% initializes the ground structure accordingly. After initializing the

% ground structure, this file executes the optimization.

function [gs] = ALTO3D_initgs(opt)

% Initialize Ground Structure Geometry

ucs = opt.ucs;

gs.b = opt.gssize(1) - 1;

gs.h = opt.gssize(2) - 1;

gs.d = opt.gssize(3) - 1;

% set up nodes and coordinates

% length of truss member on a side = 1.0

nb = gs.b+1;

nh = gs.h+1;

nd = gs.d+1;

% total # of nodes

gs.N=nb*nh*nd;

Page 360: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

331

% total # of dof

gs.Ndof = 3*gs.N;

% total # of truss members = M

% connectivity array = IEN

gs.M = gs.N*(gs.N-1)/2;

gs.IEN = zeros(2,gs.M);

member=0;

for i=1:gs.N

for j=i+1:gs.N

member=member+1;

gs.IEN(1,member)=i;

gs.IEN(2,member)=j;

end

end

% coordinate array (coordinates of node numbers)

% nodes are numbered starting at the bottom left corner and go up the

% column first and then move to the next column (see example below)

%

% *3 *6 *9 *12

%

% *2 *5 *8 *11

%

% *1 *4 *7 *10

gs.coord=zeros(3,gs.N);

number=1;

for z=0:ucs(3)/gs.d:ucs(3)

for x=0:ucs(1)/gs.b:ucs(1)

for y=0:ucs(2)/gs.h:ucs(2)

gs.coord(3,number)=z;

gs.coord(2,number)=y;

gs.coord(1,number)=x;

number=number+1;

end

end

end

% calc member lengths

gs.L = zeros(gs.M,1);

for i=1:gs.M

nodeA=gs.IEN(1,i);

nodeB=gs.IEN(2,i);

xA=gs.coord(1,nodeA);

yA=gs.coord(2,nodeA);

zA=gs.coord(3,nodeA);

xB=gs.coord(1,nodeB);

yB=gs.coord(2,nodeB);

zB=gs.coord(3,nodeB);

gs.L(i,1)=sqrt((xA-xB)^2 + (yA-yB)^2 + (zA-zB)^2);

end

% ID array (maps global node numbers to equation and DOF numbers)

gs.ID=zeros(3,gs.N);

count=1;

for i=1:gs.N

for j=1:3

Page 361: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

332

gs.ID(j,i) = count;

count = count + 1;

end

end

% member orientation

gs.theta=zeros(gs.M,3);

gs.thetab=zeros(gs.M,1);

for i=1:gs.M

nodeA=gs.IEN(1,i);

nodeB=gs.IEN(2,i);

xcoordA=gs.coord(1,nodeA);

ycoordA=gs.coord(2,nodeA);

zcoordA=gs.coord(3,nodeA);

xcoordB=gs.coord(1,nodeB);

ycoordB=gs.coord(2,nodeB);

zcoordB=gs.coord(3,nodeB);

gs.theta(i,:) = [(xcoordB - xcoordA)/gs.L(i),

(ycoordB - ycoordA)/gs.L(i),

(zcoordB - zcoordA)/gs.L(i)];

gs.thetab(i) = atan2((zcoordB-zcoordA),sqrt((xcoordB - xcoordA)^2 +

(ycoordB - ycoordA)^2));

end

% LM array (location matrix) for part 1

gs.LM=zeros(6,gs.M);

for e=1:gs.M

for a=1:2

for i=1:3

p=3*(a-1)+i;

gs.LM(p,e)=gs.ID(i,gs.IEN(a,e));

end

end

end

if any([opt.objs.type] == 6)

% Calculate midpoints of all members and calculate mid2mid distances

gs.midpts = zeros(3,gs.M);

for k = 1:gs.M

gs.midpts(:,k) = 0.5*(gs.coord(:,gs.IEN(1,k)) +

gs.coord(:,gs.IEN(2,k)));

end

% Calculate distance between all midpoints

gs.middist = 10*ones(gs.M,gs.M);

for k = 1:gs.M

for l = (k+1):gs.M

gs.middist(k,l) = norm(gs.midpts(:,l)-gs.midpts(:,k));

end

end

end

% Calculate surface element area reduction

gs.surfred = ones(1,gs.M);

for i = 1:length(gs.surfred)

coordA = gs.coord(:,gs.IEN(1,i));

coordB = gs.coord(:,gs.IEN(2,i));

surface = [coordA,coordB]==[min(gs.coord,[],2),min(gs.coord,[],2)];

surface2 = [coordA,coordB]==[max(gs.coord,[],2),max(gs.coord,[],2)];

if any(and(surface(:,1),surface(:,2))) ||

any(and(surface2(:,1),surface2(:,2)))

Page 362: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

333

sizefactor = nnz(and(surface(:,1),surface(:,2))) +

nnz(and(surface2(:,1),surface2(:,2)));

gs.surfred(i) = 1/(2*sizefactor);

continue;

end

end

% Setup initial areas and limits

gs.Ainitial = opt.init;

gs.Alower = pi/4*(opt.dlim(1)^2)*ones(1,gs.M);

gs.Aupper = pi/4*(opt.dlim(2)^2)*ones(1,gs.M);

% Initialize element area

if isempty(gs.Ainitial)

gs.Ainitial = gs.Alower;

% % Short Only

% index = gs.L <=4.25;

% gs.Ainitial(index) = gs.Aupper(1);

% 4x4x4 arrangement

%

gs.Ainitial([12,75,137,198,900,947,993,1038,1532,1563,1593,1622,1908,1923,1937,

1950])=gs.Aupper(1); % ALL X

%

gs.Ainitial([3,249,479,693,891,1073,1239,1389,1523,1641,1743,1829,1899,1953,199

1,2013])=gs.Aupper(1); % ALL Y

%

gs.Ainitial([48,111,173,234,294,353,411,468,524,579,633,686,738,789,839,888])=g

s.Aupper(1); % ALL Z

% gs.Ainitial([195,1035,1619,1947]) = gs.Aupper(1); % XY Shear 1

% gs.Ainitial([15,903,1535,1911]) = gs.Aupper(1); % XY Shear 2

% gs.Ainitial([51,297,527,741]) = gs.Aupper(1); % YZ Shear 1

% gs.Ainitial([231,465,683,885]) = gs.Aupper(1); % YZ Shear 2

% gs.Ainitial([726,777,827,876]) = gs.Aupper(1); % ZX Shear 1

% gs.Ainitial([60,123,185,246]) = gs.Aupper(1); % ZX Shear 2

% 5x5x5 arrangement

% gs.Ainitial([20,510,5142,7470,7560])=gs.Aupper(1); % ALL X

% gs.Ainitial([4,2294,5674,7454,7744])=gs.Aupper(1); % ALL Y

% gs.Ainitial([100,590,1522,2390,2800])=gs.Aupper(1); % ALL Z

% gs.Ainitial([506,3206,5281,6731,7556])=gs.Aupper(1); % ALL XY

% gs.Ainitial([104,714,1299,1859,2394])=gs.Aupper(1); % ALL YZ

% gs.Ainitial([2370,2474,2577,2679,2780])=gs.Aupper(1); % ALL ZX

% gs.Ainitial = gs.Alower + (gs.Aupper(1)-gs.Alower(1)).*rand(1,gs.M);

% gs.Ainitial = pi/4*(opt.drel^2) + (gs.Aupper(1) -

pi/4*(opt.drel^2)).*rand(1,gs.M);

end

%% 4x4x4 and 5x5x5 specifics %%

% Load appropriate overlap file

if all([opt.gssize] == [4,4,4])

% Tiers

gs.tiers = [16,32,48,64];

% load appropriate overlap file

load('4x4x4overlap.mat');

gs.overlap = overlaps;

% Set up edg_cor and faces for connectivity

Page 363: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

334

gs.edg_cor = [ 1, 4,13,16,49,52,61,64;

0, 0, 0, 0, 2,14,50,62;

0, 0, 0, 0, 3,15,51,63;

0, 0, 0, 0, 5, 8,53,56;

0, 0, 0, 0, 9,12,57,60;

0, 0, 0, 0,17,20,29,32;

0, 0, 0, 0,33,36,45,48];

gs.faces = [ 6,

7,10,11,18,19,21,24,25,28,30,31,34,35,37,40,41,44,46,47,54,55,58,59

54,55,58,59,30,31,24,21,28,25,18,19,46,47,40,37,44,41,34,35, 6,

7,10,11];

% Support neighbors Row represents the node and the values in each row

% are potential neighbor supports for that node

gs.supneighbor = [49,52,61,64;

0, 0,50,62;

0, 0,51,63;

49,52,61,64;

0, 0,53,56;

0, 0, 0,54;

0, 0, 0,55;

0, 0,53,56;

0, 0,57,60;

0, 0, 0,58;

0, 0, 0,59;

0, 0,57,60;

49,52,61,64;

0, 0,50,62;

0, 0,51,63;

49,52,61,64;

0,20,29,32;

0, 0, 0,30;

0, 0, 0,31;

0,17,29,32;

0, 0, 0,24;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0,21;

0, 0, 0,28;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0,25;

0,17,20,32;

0, 0, 0,18;

0, 0, 0,19;

0,17,20,29;

0,36,45,48;

0, 0, 0,46;

0, 0, 0,47;

0,33,45,48;

0, 0, 0,40;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0,37;

0, 0, 0,44;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0,41;

0,33,36,48;

0, 0, 0,34;

0, 0, 0,35;

Page 364: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

335

0,33,36,45;

1, 4,13,16;

0, 0, 2,14;

0, 0, 3,15;

1, 4,13,16;

0, 0, 5, 8;

0, 0, 0, 6;

0, 0, 0, 7;

0, 0, 5, 8;

0, 0, 9,12;

0, 0, 0,10;

0, 0, 0,11;

0, 0, 9,12;

1, 4,13,16;

0, 0, 2,14;

0, 0, 3,15;

1, 4,13,16];

elseif all([opt.gssize] == [5,5,5])

% Tiers

gs.tiers = [25,50,75,100,125];

% load appropriate overlap file

load('5x5x5overlap.mat');

gs.overlap = overlaps;

% Set up edg_cor and faces for connectivity

gs.edg_cor = [1,5,21,25,101,105,121,125; % Edges and corners

0,0, 0, 0, 2,102, 22,122;

0,0, 0, 0, 3,103, 23,123;

0,0, 0, 0, 4,104, 24,124;

0,0, 0, 0, 6, 10,106,110;

0,0, 0, 0, 11, 15,111,115;

0,0, 0, 0, 16, 20,116,120;

0,0, 0, 0, 26, 30, 46, 50;

0,0, 0, 0, 51, 55, 71, 75;

0,0, 0, 0, 76, 80, 96,100];

gs.faces = [ 7:9, 12:14,

17:19,27:29,31,35,36,40,41,45,47:49,52:54,56,60,61,65,66,70,72:74,77:79,81,85,8

6,90,91,95,97:99,107:109,112:114,117:119;

107:109,112:114,117:119,47:49,35,31,40,36,45,41,27:29,72:74,60,56,65,61,70,66,5

2:54,97:99,85,81,90,86,95,91,77:79, 7:9, 12:14, 17:19]';

% Support neighbors Row represents the node and the values in each row

% are potential neighbor supports for that node

gs.supneighbor = [101,105,121,125;

0, 0,102,122;

0, 0,103,123;

0, 0,104,124;

101,105,121,125;

0, 0,106,110;

0, 0, 0,107;

0, 0, 0,108;

0, 0, 0,109;

0, 0,106,110;

0, 0,111,115;

0, 0, 0,112;

0, 0, 0,113;

0, 0, 0,114;

0, 0,111,115;

Page 365: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

336

0, 0,116,120;

0, 0, 0,117;

0, 0, 0,118;

0, 0, 0,119;

0, 0,116,120;

101,105,121,125;

0, 0,102,122;

0, 0,103,123;

0, 0,104,124;

101,105,121,125;

0, 30, 46, 50;

0, 0, 0, 47;

0, 0, 0, 48;

0, 0, 0, 49;

0, 26, 46, 50;

0, 0, 0, 35;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0, 31;

0, 0, 0, 40;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0, 36;

0, 0, 0, 45;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0, 41;

0, 26, 30, 50;

0, 0, 0, 27;

0, 0, 0, 28;

0, 0, 0, 29;

0, 26, 30, 46;

0, 55, 71, 75;

0, 0, 0, 72;

0, 0, 0, 73;

0, 0, 0, 74;

0, 51, 71, 75;

0, 0, 0, 60;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0, 56;

0, 0, 0, 65;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0, 61;

0, 0, 0, 70;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0, 66;

0, 51, 55, 75;

0, 0, 0, 52;

0, 0, 0, 53;

0, 0, 0, 54;

0, 51, 55, 71;

0, 80, 96,100;

Page 366: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

337

0, 0, 0, 97;

0, 0, 0, 98;

0, 0, 0, 99;

0, 76, 96,100;

0, 0, 0, 85;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0, 81;

0, 0, 0, 90;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0, 86;

0, 0, 0, 95;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0, 0;

0, 0, 0, 91;

0, 76, 80,100;

0, 0, 0, 77;

0, 0, 0, 78;

0, 0, 0, 79;

0, 76, 80, 96;

1, 5, 21, 25;

0, 0, 2, 22;

0, 0, 3, 23;

0, 0, 4, 24;

1, 5, 21, 25;

0, 0, 6, 10;

0, 0, 0, 7;

0, 0, 0, 8;

0, 0, 0, 9;

0, 0, 6, 10;

0, 0, 11, 15;

0, 0, 0, 12;

0, 0, 0, 13;

0, 0, 0, 14;

0, 0, 11, 15;

0, 0, 16, 20;

0, 0, 0, 17;

0, 0, 0, 18;

0, 0, 0, 19;

0, 0, 16, 20;

1, 5, 21, 25;

0, 0, 2, 22;

0, 0, 3, 23;

0, 0, 4, 24;

1, 5, 21, 25];

else

warning('Non-standard opt.gssize , several options may be unavailable')

end

gs.removednode = false(gs.M,1);

if ~isempty(opt.removenode)

for i = 1:gs.M

if any(any(gs.IEN(:,i) == [opt.removenode]))

gs.removednode(i,1) = true;

end

end

Page 367: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

338

end

end

Overlap Function

The overlap function is a standalone function that is run prior to the optimization

problem. It can be used to generate the overlapping elements matrix for different size cubic arrays

of elements. In general, it was run once for 4x4x4 and 5x5x5 node arrays. Running the function

creates a file that is read by the ground structure initialization function for the overlap penalty.

ALTO_overlap.m

% This code generates the binary overlap matrix offline so that it does

% need to be calculated in real time. The same 4x4x4 or 5x5x5 binary matrix

% applies for every UCS as long as the full ground structure is considered.

% Most of the options included below are arbitrary but required for

% ALTO3D_initgs.m to run

% The overlaps matrix should be saved to a file that can be loaded in

% ALTO3D_initgs.m

close all;

clear;

clc;

array = 5; % Array Size (5x5x5 or 4x4x4, etc)

opt.ucs = 2; % Unit cell size in mm

opt.diam = 1; % Beam Diameter

opt.mod = 2e5; % Modulus of elasticity N/mm^2

opt.dlim = [9.5e-5, 9.5]; % Element diameter limits [min,max] in mm

opt.drel = [0.3]; % Relevant diameter (mm) - Used to determine which

members are relevant

opt.gssize = [ array, array, array];

opt.init = pi/4*(opt.dlim(1)^2)*ones(1,((array^3)*(array^3-1)/2));

gs = ALTO3D_initgs(opt);

overlaps = false(gs.M,gs.M);

pairs = [];

solutions = [];

for i = 1:gs.M

for j = (i+1):gs.M

ptL1 = gs.coord(:,gs.IEN(1,i)); % L1 starting point

ptL2 = gs.coord(:,gs.IEN(1,j)); % L2 starting point

L1 = gs.L(i); % L1 length

L2 = gs.L(j); % L2 length

dir1 = gs.theta(i,:)'; % L1 direction unit vector

dir2 = gs.theta(j,:)'; % L2 direction unit vector

% Ax = b, solve for positions along lines where they intersect

b = ptL2 - ptL1;

Page 368: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

339

A = [L1*dir1(1) -L2*dir2(1)

L1*dir1(2) -L2*dir2(2)

L1*dir1(3) -L2*dir2(3)];

options = optimoptions(@lsqlin,'Display','None');

x = lsqlin(A,b,[],[],[],[],[0.01,0.01],[0.99,0.99],[],options);

err = b-A*x;

% Check if result is an intersection point

if (all(abs(err) < .001))

% disp(['Found overlap',num2str(size(solutions,2)+1)])

overlaps(i,j) = true;

pairs = [pairs,[gs.IEN(:,i);gs.IEN(:,j)]];

solutions = [solutions,x];

else

continue

end

end

end

%% This section shows an example of how to evaluate whether the relevant

% members of the current state are overlapping.

A = gs.Ainitial;

A([8,20,24,55,67,75,104,106,108,110,153,161,192,196,217,219,227,250,256,275,281

,302,323,334]) = 1;

ALTO3D_plot(A,gs.IEN,gs.coord,.1)

rel_overlaps = overlaps((A>.1),(A>.1));

nnz(rel_overlaps)

Unit Cell Simulation Function

The unit cell simulation file is the function given to the optimization algorithm by the

executive function to simulate a unit cell design. It contains calls to any of the objective functions

and code to evaluate penalty functions for a unit cell design. The result of this file is the objective

function values for the optimization algorithm.

ALTO3D_ucsim.m

function [obj,constrs] = ALTO3D_ucsim(Areduced,opt,optim,gs)

% This function executes the appropriate simulations and combines the

% resulting objective functions to a single objective function value.

objtypes = [opt.objs.type];

Ain = Areduced;

Page 369: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

340

% Rebuild A

Areduced([optim.keep]) = Areduced;

Areduced([optim.remove]) = gs.Alower(1);

A = zeros(1,gs.M);

if isfield(optim,'en_symmap')

for i = 1:gs.M

A(i) = Areduced(optim.en_symmap(i));

end

end

A(gs.removednode) = gs.Alower(1);

A_relevant = (pi/4*(opt.drel)^2);

A_rel = A>=A_relevant;

Azero = (A < A_relevant);

A(Azero) = gs.Alower(1);

% Reduce areas for shared surface elements (ONLY USE FOR HOMOGENIZATION

% PROPERTIES)

A_surfred = A;

A_surfred(A_rel) = gs.surfred(A_rel).*(A(A_rel));

score = zeros(1,length(objtypes));

objs = zeros(1,length(objtypes));

% Apply constraint if borg

switch lower(opt.algorithm)

case{'borg'}

if ~isempty(optim.Vstar)

constrs = (A_surfred*gs.L) - optim.Vstar;

% constrs = Ain*(optim.L.*optim.en_count) - optim.Vstar;

pass = (constrs<=0);

constrs(pass) = 0;

if constrs ~= 0

obj = 1e15*ones(1,length(objtypes));

obj(3) = abs((A_surfred*gs.L)/((opt.ucs)^3) -

optim.volume.vftarget)/optim.volume.vftarget;

% constrs = 0;

return

end

else

constrs = 0;

end

end

switch lower(opt.algorithm)

case{'borg','nsga-ii'}

i = 1;

while (i <= length(objtypes))

if objtypes(i) == 1

if (i+1 <= length(objtypes))

if objtypes(i+1) == 2

opt.type = 12;

objs(i:(i+1)) = ALTO3D_compltherm(A,opt,optim,gs)';

i = i + 2;

else

opt.type = 1;

objs(i) = nonzeros(ALTO3D_compltherm(A,opt,optim,gs));

i = i + 1;

Page 370: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

341

end

else

opt.type = 1;

objs(i) = nonzeros(ALTO3D_compltherm(A,opt,optim,gs));

i = i + 1;

end

elseif objtypes(i) == 2

if (i+1 <= length(objtypes))

if objtypes(i+1) == 1

opt.type = 12;

objs(i:(i+1)) = ALTO3D_compltherm(A,opt,optim,gs)';

i = i + 2;

else

opt.type = 2;

objs(i) = nonzeros(ALTO3D_compltherm(A,opt,optim,gs));

i = i + 1;

end

else

opt.type = 2;

objs(i) = nonzeros(ALTO3D_compltherm(A,opt,optim,gs));

i = i + 1;

end

elseif objtypes(i) == 3

if all([opt.gssize]==[4,4,4])

Eh = ALTO3D_homog(A_surfred,A,opt,optim,gs);

elseif all([opt.gssize]==[5,5,5])

Eh = ALTO3D_homog5(A_surfred,A,opt,optim,gs);

else

error('Homogenization cannot be performed with this array

size')

end

% objs(i) = sqrt(sumsqr((optim.homog.target -

Eh)./optim.homog.target)); % Normalized

% objs(i) = sqrt(sumsqr((optim.homog.target - Eh))); % NOT

normalized

a = abs(((optim.homog.target - Eh)./optim.homog.target));

objs(i) = sum(a(a<inf));

i = i + 1;

elseif objtypes(i) == 4

if optim.volume.vftarget > 0

objs(i) = abs((A_surfred*gs.L)/((opt.ucs)^3) -

optim.volume.vftarget)/optim.volume.vftarget;

i = i + 1;

elseif abs(optim.volume.vftarget) <= 1e-10

objs(i) = A_surfred*gs.L;

i = i + 1;

end

elseif objtypes(i) == 5

objs(i) = nnz(A_rel);

i = i + 1;

elseif objtypes(i) == 6

A_red = A(A_rel); %

r_rel = (A_red/pi).^(.5);

middist_red = gs.middist((A_rel),(A_rel));

middist_adj = middist_red;

for k = 1:length(A_red)

for l = (k+1):length(A_red)

middist_adj(k,l) = middist_red(k,l)-

(r_rel(k)+r_rel(l));

end

end

Page 371: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

342

if optim.powder.type == 1

objs(i) = mean(mean(middist_adj(middist_adj<900)));

elseif optim.powder.type == 2

objs(i) = min(min(middist_adj(middist_adj<900)));

elseif optim.powder.type == 3

objs(i) = nnz(middist_adj < optim.powder.tol);

end

i = i + 1;

end

end

% Penalties

P = [0];

if opt.Ptype == 1

elseif opt.Ptype == 2

% Overlap Penalty

rel_overlaps = gs.overlap((A_rel),(A_rel));

if nnz(rel_overlaps) > 0

P = [P,100];

end

elseif opt.Ptype == 3

% Connectivity

rel_nodes = unique(gs.IEN(:,(A_rel)'));

check1 = sum(ismember(gs.edg_cor,rel_nodes),2);

check2 = sum(ismember(gs.faces,rel_nodes),2);

if any([check1;check2]==1)

flaws = nnz([check1;check2]==1);

P = [P,100];

end

else

% Overlap Penalty

rel_overlaps = gs.overlap((A_rel),(A_rel));

if nnz(rel_overlaps) > 0

P = [P,100];

end

% Connectivity

rel_nodes = unique(gs.IEN(:,(A_rel)'));

check1 = sum(ismember(gs.edg_cor,rel_nodes),2);

check2 = sum(ismember(gs.faces,rel_nodes),2);

if any([check1;check2]==1)

flaws = nnz([check1;check2]==1);

P = [P,100];

end

end

% Support Structure Penalties

if any(opt.SPtype == [2,3])

gs.IENred = gs.IEN(:,A_rel);

horz = gs.thetab(A_rel) == 0;

Page 372: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

343

if all([opt.gssize] == [4,4,4])

t1 = unique(gs.IENred(gs.IENred<=gs.tiers(1)))';

grounded = [];

neighbored = [];

new = unique(gs.IENred(2,(ismember(gs.IENred(1,:),t1) &

gs.IENred(2,:) > gs.tiers(1))));

if opt.SPtype == 2

for i = new

neighbored = [neighbored,gs.supneighbor(i,:)];

end

end

grounded = unique([grounded,new,neighbored(neighbored >0)]);

t2 = unique(grounded(grounded>gs.tiers(1) &

grounded<=gs.tiers(2)));

new = unique(gs.IENred(2,(ismember(gs.IENred(1,:),t2) &

gs.IENred(2,:) > gs.tiers(2))));

grounded = [grounded,new];

if opt.SPtype == 2

for i = new

neighbored = [neighbored,gs.supneighbor(i,:)];

end

end

grounded = unique([grounded,neighbored(neighbored >0)]);

t3 = unique(grounded(grounded>gs.tiers(2) &

grounded<=gs.tiers(3)));

new = unique(gs.IENred(2,(ismember(gs.IENred(1,:),t3) &

gs.IENred(2,:) > gs.tiers(3))));

grounded = [grounded,new];

if opt.SPtype == 2

for i = new

neighbored = [neighbored,gs.supneighbor(i,:)];

end

end

grounded = unique([grounded,neighbored(neighbored >0)]);

t4 = unique(grounded(grounded>gs.tiers(3) &

grounded<=gs.tiers(4)));

elseif all([opt.gssize] == [5,5,5])

t1 = unique(gs.IENred(gs.IENred<=gs.tiers(1)))';

grounded = [];

neighbored = [];

new = unique(gs.IENred(2,(ismember(gs.IENred(1,:),t1) &

gs.IENred(2,:) > gs.tiers(1))));

if opt.SPtype == 2

for i = new

neighbored = [neighbored,gs.supneighbor(i,:)];

end

end

grounded = unique([grounded,new,neighbored(neighbored >0)]);

t2 = unique(grounded(grounded>gs.tiers(1) &

grounded<=gs.tiers(2)));

new = unique(gs.IENred(2,(ismember(gs.IENred(1,:),t2) &

gs.IENred(2,:) > gs.tiers(2))));

grounded = [grounded,new];

if opt.SPtype == 2

for i = new

neighbored = [neighbored,gs.supneighbor(i,:)];

end

Page 373: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

344

end

grounded = unique([grounded,neighbored(neighbored >0)]);

t3 = unique(grounded(grounded>gs.tiers(2) &

grounded<=gs.tiers(3)));

new = unique(gs.IENred(2,(ismember(gs.IENred(1,:),t3) &

gs.IENred(2,:) > gs.tiers(3))));

grounded = [grounded,new];

if opt.SPtype == 2

for i = new

neighbored = [neighbored,gs.supneighbor(i,:)];

end

end

grounded = unique([grounded,neighbored(neighbored >0)]);

t4 = unique(grounded(grounded>gs.tiers(3) &

grounded<=gs.tiers(4)));

new = unique(gs.IENred(2,(ismember(gs.IENred(1,:),t4) &

gs.IENred(2,:) > gs.tiers(4))));

grounded = [grounded,new];

if opt.SPtype == 2

for i = new

neighbored = [neighbored,gs.supneighbor(i,:)];

end

end

grounded = unique([grounded,neighbored(neighbored >0)]);

t5 = unique(grounded(grounded>gs.tiers(4) &

grounded<=gs.tiers(5)));

end

if opt.SPtype == 2 % Neighbored

horz = gs.thetab(A_rel) == 0;

flaws = nnz(~ismember(gs.IENred(1,:),grounded)) +

nnz(~ismember(gs.IENred(2,horz),grounded));

elseif opt.SPtype == 3

if all([opt.gssize] == [4,4,4])

grounded = unique([grounded,t4-48]);

elseif all([opt.gssize] == [5,5,5])

grounded = unique([grounded,t5-100]);

end

horz = gs.thetab(A_rel) == 0;

flaws = nnz(~ismember(gs.IENred(1,:),grounded)) +

nnz(~ismember(gs.IENred(2,horz),grounded));

else

error('Unknown SPtype')

end

if flaws > 0

P = [P,100];

end

end

obj = sum(P)*ones(1,length(objtypes)) +

([opt.objs.minimize].*[opt.objs.scale].*[optim.wght]).*objs;

case{'fmincon'}

i = 1;

while (i <= length(objtypes))

if objtypes(i) == 1

Smin = optim.compl.range(1);

Smax = optim.compl.range(2);

if (i+1 <= length(objtypes))

Page 374: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

345

if objtypes(i+1) == 2

opt.type = 12;

Condmin = optim.therm.range(1);

Condmax = optim.therm.range(2);

score(i:(i+1)) = ALTO3D_compltherm(A,opt,optim,gs)';

objs(i:(i+1)) = [(Smin - score(i))/(Smin - Smax),

(Condmin - score(i+1))/(Condmin - Condmax)];

i = i + 2;

end

else

opt.type = 1;

score(i) = nonzeros(ALTO3D_compltherm(A,opt,optim,gs));

% objs(i) = (Smin - score(i))/(Smin - Smax);

objs(i) = score(i);

i = i + 1;

end

elseif objtypes(i) == 2

Condmin = optim.therm.range(1);

Condmax = optim.therm.range(2);

if (i+1 <= length(objtypes))

if objtypes(i+1) == 1

Smin = optim.compl.range(1);

Smax = optim.compl.range(2);

opt.type = 12;

score(i:(i+1)) = ALTO3D_compltherm(A,opt,optim,gs)';

objs(i:(i+1)) = [(Smin - score(i))/(Smin - Smax),

(Condmin - score(i+1))/(Condmin - Condmax)];

i = i + 2;

end

else

opt.type = 2;

score(i) = nonzeros(ALTO3D_compltherm(A,opt,optim,gs));

objs(i) = (Condmin - score(i))/(Condmin - Condmax);

i = i + 1;

end

elseif objtypes(i) == 3

Eh = ALTO3D_homog(A_surfred,opt,optim,gs);

% objs(i) = sqrt(sumsqr((optim.homog.target -

Eh)./optim.homog.target)); % Normalized

% objs(i) = sqrt(sumsqr((optim.homog.target - Eh))); % NOT

normalized

a = abs(((optim.homog.target - Eh)./optim.homog.target));

objs(i) = sum(a(a<inf),'all');

i = i + 1;

elseif objtypes(i) == 4

target = (optim.volume.vftarget)*optim.volume.range(2);

score(i) = A*gs.L;

% objs(i) = abs(target - score(i))/target; % Normalized

% objs(i) = 1e6*abs(target - score(i))/target; % NOT Normalized

% objs(i) = abs(target - score(i))/optim.volume.range(2); %

Zero volume

objs(i) = score(i)/optim.volume.range(2);

i = i + 1;

end

end

% Calculate Penalty functions

P = 1;

if ~isempty(opt.drel)

if ~isempty(opt.Pcount)

Page 375: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

346

ne_rel = nnz(A >= A_relevant);

P = [P,(ne_rel/gs.M)^(opt.Pcount)];

end

if ~isempty(opt.Pdmid)

P2 = ones(1,length(A));

num = gs.M;

for i = 1:length(A)

if A(i) > A_relevant/10

P2(i) = (A(i)/A_relevant);

else

P2(i) = .1;

end

end

P2 = prod(P2)^(opt.Pdmid/num);

P = [P,P2];

end

end

obj =

prod(P)*([opt.objs.minimize].*[opt.objs.scale].*[optim.wght])*objs';

if (opt.saveobj == 1)

load(opt.tempname,'outvars')

iter = size(outvars,2);

if ~isempty(outvars(iter).vol)

iter = iter + 1;

outvars(iter).Iteration = iter;

end

outvars(iter).objs = [[outvars(iter).objs];objs];

outvars(iter).P = [[outvars(iter).P];P];

save(opt.tempname,'outvars','-append')

end

end

end

Compliance and Thermal Conduction Function

The compliance and thermal conduction function computes the strain energy objective

function and the thermal conduction objective function.

ALTO3D_compltherm.m

function [objs] = ALTO3D_compltherm(A,opt,optim,gs)

del_strain = 1;

L = gs.L;

L_orig = gs.L;

objs = [0;0];

if (opt.type == 2 || opt.type == 12)

in = optim.therm.in;

out = optim.therm.out;

int = [];

for i = 1:gs.M

for j = 1:2

Page 376: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

347

n = gs.IEN(j,i);

if ~any(n==[in,out,int])

int = [int,n];

end

end

end

countint = length(int);

num_nodes = max([int,in,out]);

rmindex = zeros(gs.M,1);

% if ~isempty(opt.removelength)

% rmindex = rmindex | ((gs.L > opt.removelength(1) & gs.L <

opt.removelength(2)));

% end

% if ~isempty(opt.buildang)

% if isfield(optim,'thetab')

% rmindex = rmindex | ((abs(gs.thetab) > pi*opt.buildang(1)/180) &

(abs(gs.thetab) < pi*opt.buildang(2)/180));

% end

% end

% Try this instead to remove any zeroed elements

if ~all(A'==gs.Alower(1))

rmindex = A'==gs.Alower(1);

end

num_eqs = countint + nnz(~rmindex);

Tfixed = zeros(num_nodes,1);

Tfixed(in) = optim.therm.T1;

Tfixed(out) = optim.therm.T2;

% map interior node number to equation number

for i = 1:countint

nn(int(i)) = i;

end

end

iter = 1;

% For coupled thermo mechanical response, adjust number of iterations to allow

iterative solution

% DO NOT USE coupled thermo mechanical response with heat source (current

source) input

while ((abs(del_strain(end)) > 1e-4) && (iter < 2))

fa = zeros(gs.Ndof,1);

if (opt.type == 2 || opt.type == 12)

% Mem contains all the member information for the thermal circuit

% [ Node1, Node2, Length, Area, Resistance]

Mem(:,1:5) = [gs.IEN',L,A',L./(opt.kcond*A')];

Mem_red = Mem(~rmindex,:);

% Initialize b vector and K matrix for a = K\b

b = zeros(num_eqs,1);

if ~isempty(opt.Iin)

b(end-(length(opt.Iin)-1):end) = opt.Iin;

end

K = zeros(num_eqs); % N2 N3 i1 i2 i3 i4 i5

Page 377: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

348

% Populate resistor network matrix

for i = 1:nnz(~rmindex)

n1 = Mem_red(i,1);

n2 = Mem_red(i,2);

if ismember(n1,[in,out]) && ismember(n2,[in,out])

b(i) = Tfixed(n1) - Tfixed(n2);

elseif ismember(n1,[in,out]) && ismember(n2,int)

b(i) = Tfixed(n1);

K(i,nn(n2)) = 1;

elseif ismember(n1,int) && ismember(n2,int)

K(i,nn([n1,n2])) = [-1,1];

elseif ismember(n1,int) && ismember(n2,[in,out])

b(i) = -Tfixed(n2);

K(i,nn(n1)) = -1;

else

disp('Warning, unexpected node locations')

disp(['Node 1 = ',num2str(n1),' Node 2 = ',num2str(n2)])

end

for j = 1:countint

if n1 == int(j)

K(nnz(~rmindex)+j,countint+i) = 1;

end

if n2 == int(j)

K(nnz(~rmindex)+j,countint+i) = -1;

end

end

K(i,(countint+i)) = Mem_red(i,5);

end

scale = 1;

Ks = sparse(K);

% Determines how to treat singular matrix

if condest(Ks) == inf

if opt.type == 2

objs(:,iter) = [0;-1e20];

else

objs(:,iter) = [1e20;-1e20];

end

return

else

a = (1/scale)*(Ks\(scale*b));

end

[warnmsg, msgid] = lastwarn;

% Determines how to treat singular or near singular matrix

if strcmp(msgid,'MATLAB:singularMatrix') || strcmp(msgid,

'MATLAB:nearlySingularMatrix')

if opt.type == 2

objs(:,iter) = [0;-1e20];

else

objs(:,iter) = [1e20;-1e20];

end

lastwarn('')

return

end

% Populate Vector of Voltages at each Node

V = zeros(num_nodes,1);

Page 378: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

349

for i=1:length(nn)

if nn(i) ~= 0

V(i) = a(nn(i));

end

end

V = V + Tfixed;

% Increase size of Mem to be N1, N2, Length, Area, Resistance, Temp

Drop,

% Temp Avg, Current

for i = 1:nnz(~rmindex)

Mem_red(i,6) = V(Mem_red(i,1)) - V(Mem_red(i,2));

Mem_red(i,7) = (V(Mem_red(i,1)) + V(Mem_red(i,2)))/2;

end

Mem_red(:,8) = a(countint+1:end);

Mem(~rmindex,6:8) = Mem_red(:,6:8);

total_i = 0;

for i = 1:nnz(~rmindex)

% if ismember(Mem_red(i,1),[in,int]) ||

ismember(Mem_red(i,2),[in,int])

if ismember(Mem_red(i,1),[out]) || ismember(Mem_red(i,2),[out])

total_i = total_i + abs(a(countint+i));

end

end

% Calculation of effective conductivity with in and out voltages

% Reff(iter) = (optim.therm.T1-optim.therm.T2)/(total_i);

% objs(2,iter) = 1/Reff(iter);

% Calculation of effective conductivity with input current

Reff(iter) = (mean(V)/(total_i));

objs(2,iter) = 1/Reff(iter);

%% For coupled thermo mechanical response, uncomment this section

% % Thermal Applied Forces

% fa_th = zeros(gs.Ndof,1);

% for i = 1:gs.M

% fth = opt.mod*Mem(i,4)*opt.alpha*(Mem(i,7)-optim.therm.Tref);

% dof_i = gs.LM(:,i);

% fa_th(dof_i) = fa_th(dof_i) + [-fth*gs.theta(i,:)';

% fth*gs.theta(i,:)'];

% end

% fa = fa + fa_th;

% if iter > 1

% del_Reff(iter) = (Reff(end) - Reff(end-1))/Reff(end-1)*100;

% end

end

if (opt.type == 1 || opt.type == 12)

% Applied Forces

for i = 1:size(optim.compl.loads,1)

node = optim.compl.loads(i,1);

fa((3*node-2):3*node) = fa(3*node-2:3*node) +

optim.compl.loads(i,2:4)';

end

Page 379: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

350

% Fixed BC's

bcf = [];

for i = 1:size(optim.compl.bcs,1)

for j = 1:3

if optim.compl.bcs(i,j+1) == 0

bcf = [bcf,3*(optim.compl.bcs(i,1)-1) + j];

end

end

end

% local member stiffness = AE/L

% element stiffness matrix = k

% global stiffness matrix for part 1 = K1

K1=zeros(gs.Ndof,gs.Ndof);

dk1v=zeros(36,gs.M);

for i=1:gs.M

Cx = gs.theta(i,1);

Cy = gs.theta(i,2);

Cz = gs.theta(i,3);

lambda = [ Cx*Cx Cx*Cy Cx*Cz;

Cx*Cy Cy*Cy Cy*Cz;

Cx*Cz Cy*Cz Cz*Cz];

k = [ lambda, -lambda;

-lambda, lambda];

dk = opt.mod/L_orig(i)*k;

k = A(i)*opt.mod/L_orig(i)*k;

dk1v(:,i) = reshape(dk',[36,1]);

q = gs.LM(:,i);

K1(q,q) = K1(q,q) + k;

end

index = true(1,gs.Ndof);

index(bcf) = false;

nn2 = 1:gs.Ndof;

nn2_r = nn2(index);

fa_r = fa(index);

K1_r = K1(index,:);

K1_r = K1_r(:,index);

% solve for ua

ua_r = K1_r\fa_r;

strain(iter) = ua_r'*K1_r*ua_r;

objs(1,iter) = strain(iter);

% Update member lengths

ua = zeros(gs.Ndof,1);

ua(nn2_r) = ua_r;

ua_xy = reshape(ua,[3,gs.Ndof/3]);

coord_disp = gs.coord + ua_xy;

for i=1:gs.M

nodeA=gs.IEN(1,i);

nodeB=gs.IEN(2,i);

xA=coord_disp(1,nodeA);

yA=coord_disp(2,nodeA);

Page 380: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

351

zA=coord_disp(3,nodeA);

xB=coord_disp(1,nodeB);

yB=coord_disp(2,nodeB);

zB=coord_disp(3,nodeB);

L(i,1)=sqrt((xA-xB)^2+(yA-yB)^2+(zA-zB)^2);

end

% % Plot deformed truss

% ALTO3D_plotdeformed(A,gs,coord_disp)

if iter > 1

del_strain(iter) = abs((strain(end) - strain(end-1))/strain(end-

1))*100;

end

end

if (opt.type == 1 || opt.type == 2)

break

end

iter = iter + 1;

end

objs = objs(:,end);

end

Homogenization Functions

Three different homogenization functions are included. The first “ALTO3D_homog.m”

is for a 4x4x4 nodal array, the second is for a 5x5x5 nodal array “ALTO3D_homog5.m”, and the

third is for updated boundary conditions, “ALTO3D_homog5v2.m”, that represent the relaxed

modification discussed in Chapter 5. In all cases, the homogenization function is called by the

unit cell simulation function and computes the homogenized constitutive matrix.

ALTO3D_homog.m

function [Eh,min2buckle] = ALTO3D_homog(A,Abuck,opt,optim,gs)

A_relevant = (pi/4*(opt.drel)^2);

A_rel = A>=A_relevant;

% local member stiffness = AE/L

% element stiffness matrix = k

% global stiffness matrix for part 1 = K1

K1=zeros(gs.Ndof,gs.Ndof);

k_all=zeros(6,6*gs.M);

dk1v=zeros(36,gs.M);

for i=1:gs.M

Cx = gs.theta(i,1);

Cy = gs.theta(i,2);

Page 381: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

352

Cz = gs.theta(i,3);

lambda = [ Cx*Cx Cx*Cy Cx*Cz;

Cx*Cy Cy*Cy Cy*Cz;

Cx*Cz Cy*Cz Cz*Cz];

k = [ lambda, -lambda;

-lambda, lambda];

dk = opt.mod/gs.L(i)*k;

k = A(i)*opt.mod/gs.L(i)*k;

dk1v(:,i) = reshape(dk',[36,1]);

k_all(1:6,(6*i-5):(6*i)) = k;

q = gs.LM(:,i);

K1(q,q) = K1(q,q) + k;

end

%% Modify stiffness matrix for periodic BC's for each strain test

%%% Unit test strain cases 1 and 2 (axial strain), 3 (shear strain)

K1_1 = K1;

%%%% Boundary Conditons %%%%

% This code is for reducing K1 using the following equations where the

% numbers are the corresponding degree of freedom. These equations come

% from the following general periodicity equations (van der Sluis 2000):

% u_11 = u_12 - u_v4

% u_22 = u_21 - u_v1

% u_v3 = u_v2 + u_v4 - u_v1

%

% Using the general form

% a = b - c + d -or- a = b - c + d + e1 - e2

% Symmetry face nodes % CAREFUL WHICH NODES ARE CALCULATED FIRST, 14/15

% must be calculated before 2/3 can be calculated

xnodes = [18,19,20,34,35,36,50,51;

30,31,32,46,47,48,62,63;

13,13,13,13,13,13,13,13;

1, 1, 1, 1, 1, 1, 1, 1];

ynodes = [21,25,29,37,41,45,53,57;

24,28,32,40,44,48,56,60;

4, 4, 4, 4, 4, 4, 4, 4;

1, 1, 1, 1, 1, 1, 1, 1];

znodes = [ 6, 7, 8,10,11,12,14,15;

54,55,56,58,59,60,62,63;

49,49,49,49,49,49,49,49;

1, 1, 1, 1, 1, 1, 1, 1];

others = [ 2, 3, 5, 9,17,33;

14,15,53,57,20,36;

13,13,49,49, 4, 4;

1, 1, 1, 1, 1, 1];

nodes = [ xnodes, ynodes, znodes, others];

a1 = zeros(1,30);

b1 = zeros(1,30);

Page 382: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

353

c1 = zeros(1,30);

d1 = zeros(1,30);

for i=1:length(nodes(1,:))

a1((3*i-2):3*i) = (3*nodes(1,i)-2):3*nodes(1,i);

b1((3*i-2):3*i) = (3*nodes(2,i)-2):3*nodes(2,i);

c1((3*i-2):3*i) = (3*nodes(3,i)-2):3*nodes(3,i);

d1((3*i-2):3*i) = (3*nodes(4,i)-2):3*nodes(4,i);

end

for i = 1:length(a1)

K1_1(:,b1(i)) = K1_1(:,b1(i)) + K1_1(:,a1(i));

K1_1(:,c1(i)) = K1_1(:,c1(i)) - K1_1(:,a1(i));

K1_1(:,d1(i)) = K1_1(:,d1(i)) + K1_1(:,a1(i));

end

% for i = 1:length(a1)

% K1_1(b1(i),:) = K1_1(b1(i),:) + K1_1(a1(i),:);

% end

K1_2 = K1_1;

K1_3 = K1_1;

K1_4 = K1_1;

K1_5 = K1_1;

K1_6 = K1_1;

% % Prescribed Displacements

pd1 = zeros(gs.Ndof,1);

pd2 = zeros(gs.Ndof,1);

pd3 = zeros(gs.Ndof,1);

pd4 = zeros(gs.Ndof,1);

pd5 = zeros(gs.Ndof,1);

pd6 = zeros(gs.Ndof,1);

nn = (1:gs.Ndof)';

INDEX = true(6,gs.Ndof); % All dof that will be removed

INDEX(:,a1) = false; % Set periodic dofs to be removed

PRDOF = zeros(6,gs.Ndof); % All dof with strain

% Strain (e); Free (f); Fixed (0);

e = 1;

f = 2;

%% Test 1

% Prescribed BC's

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,13) = [e;0;0]; % Strain in x, else fix

prenodes(:,[1,4,49]) = repmat([0;0;0],[1,3]); % All fixed

prenodes(:,[24,28,40,44]) = repmat([f;0;f],[1,4]); % Fix y

prenodes(:,[30,31,46,47]) = repmat([e;f;f],[1,4]); % Strain x

prenodes(:,[54,55,58,59]) = repmat([f;f;0],[1,4]); % Fix z

prenodes(:,[32,48]) = repmat([e;0;f],[1,2]); % Strain x, fix y

prenodes(:,[56,60]) = repmat([f;0;0],[1,2]); % Fix y,z

prenodes(:,[62,63]) = repmat([e;f;0],[1,2]); % Strain x, fix z

% Calculated Corners

prenodes(:,16) = prenodes(:,13) + prenodes(:, 4) - prenodes(:, 1);

prenodes(:,52) = prenodes(:,49) + prenodes(:, 4) - prenodes(:, 1);

prenodes(:,61) = prenodes(:,49) + prenodes(:,13) - prenodes(:, 1);

prenodes(:,64) = prenodes(:,49) + prenodes(:,13) + prenodes(:, 4) -

2*prenodes(:,1);

Page 383: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

354

% Generate PRDOF showing strains and index for removing prescribed dofs

for i = 1:size(prenodes,2)

for j = 1:3

dof = (3*(i-1) + j);

bc = prenodes(j,i);

if (bc == 0 || bc == e)

INDEX(1,dof) = false;

PRDOF(1,dof) = bc;

end

end

end

%Prepare Force Vector

F1 = -K1_1*(PRDOF(1,:)');

% for i = 1:length(a1)

% F1(b1(i)) = F1(b1(i)) + F1(a1(i));

% end

% Reduce vectors and matrices based on prescribed and periodic

nn1 = nn(INDEX(1,:));

F1_r = F1(INDEX(1,:)); % Reduced Force vector

K1_1r = K1_1(INDEX(1,:),:);

K1_1r = K1_1r(:,INDEX(1,:));

%% Test 2

% Prescribed BC's

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,4) = [0;e;0]; % Strain in y, else fix

prenodes(:,[1,13,49]) = repmat([0;0;0],[1,3]); % Fix all

prenodes(:,[24,28,40,44]) = repmat([f;e;f],[1,4]); % Free x,z Strain y

prenodes(:,[30,31,46,47]) = repmat([0;f;f],[1,4]); % Fix x Free y,z

prenodes(:,[54,55,58,59]) = repmat([f;f;0],[1,4]); % Free x,y Fix z

prenodes(:,[32,48]) = repmat([0;e;f],[1,2]); % Fix x Strain y Free z

prenodes(:,[56,60]) = repmat([f;e;0],[1,2]); % Free x Strain y Fix z

prenodes(:,[62,63]) = repmat([0;f;0],[1,2]); % Fix x Free y Fix z

% Calculated Corners

prenodes(:,16) = prenodes(:,13) + prenodes(:, 4) - prenodes(:, 1);

prenodes(:,52) = prenodes(:,49) + prenodes(:, 4) - prenodes(:, 1);

prenodes(:,61) = prenodes(:,49) + prenodes(:,13) - prenodes(:, 1);

prenodes(:,64) = prenodes(:,49) + prenodes(:,13) + prenodes(:, 4) -

2*prenodes(:,1);

% Generate PRDOF showing strains and index for removing prescribed dofs

for i = 1:size(prenodes,2)

for j = 1:3

dof = (3*(i-1) + j);

bc = prenodes(j,i);

if (bc == 0 || bc == e)

INDEX(2,dof) = false;

PRDOF(2,dof) = bc;

end

end

end

%Prepare Force Vector

F2 = -K1_2*(PRDOF(2,:)');

% for i = 1:length(a1)

% F2(b1(i)) = F2(b1(i)) + F2(a1(i));

% end

Page 384: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

355

% Reduce vectors and matrices based on prescribed and periodic

nn2 = nn(INDEX(2,:));

F2_r = F2(INDEX(2,:)); % Reduced Force vector

K1_2r = K1_2(INDEX(2,:),:);

K1_2r = K1_2r(:,INDEX(2,:));

%% Test 3

% Prescribed BC's

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,49) = [0;0;e]; % Strain in x, else fix

prenodes(:,[1,4,13]) = repmat([0;0;0],[1,3]); % All fixed

prenodes(:,[24,28,40,44]) = repmat([f;0;f],[1,4]); % Fix y

prenodes(:,[30,31,46,47]) = repmat([0;f;f],[1,4]); % Strain x

prenodes(:,[54,55,58,59]) = repmat([f;f;e],[1,4]); % Fix z

prenodes(:,[32,48]) = repmat([0;0;f],[1,2]); % Strain x, fix y

prenodes(:,[56,60]) = repmat([f;0;e],[1,2]); % Fix y,z

prenodes(:,[62,63]) = repmat([0;f;e],[1,2]); % Strain x, fix z

% Calculated Corners

prenodes(:,16) = prenodes(:,13) + prenodes(:, 4) - prenodes(:, 1);

prenodes(:,52) = prenodes(:,49) + prenodes(:, 4) - prenodes(:, 1);

prenodes(:,61) = prenodes(:,49) + prenodes(:,13) - prenodes(:, 1);

prenodes(:,64) = prenodes(:,49) + prenodes(:,13) + prenodes(:, 4) -

2*prenodes(:,1);

% Generate PRDOF showing strains and index for removing prescribed dofs

for i = 1:size(prenodes,2)

for j = 1:3

dof = (3*(i-1) + j);

bc = prenodes(j,i);

if (bc == 0 || bc == e)

INDEX(3,dof) = false;

PRDOF(3,dof) = bc;

end

end

end

%Prepare Force Vector

F3 = -K1_3*(PRDOF(3,:)');

% for i = 1:length(a1)

% F3(b1(i)) = F3(b1(i)) + F3(a1(i));

% end

% Reduce vectors and matrices based on prescribed and periodic

nn3 = nn(INDEX(3,:));

F3_r = F3(INDEX(3,:)); % Reduced Force vector

K1_3r = K1_3(INDEX(3,:),:);

K1_3r = K1_3r(:,INDEX(3,:));

%% Test 4

% Prescribed BC's

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,4) = [0;0;e]; % Strain in x, else fix

prenodes(:,[1,13,49]) = repmat([0;0;0],[1,3]); % All fixed

prenodes(:,[24,28,40,44]) = repmat([f;0;e],[1,4]); % Fix y

prenodes(:,[30,31,46,47]) = repmat([0;f;f],[1,4]); % Strain x

prenodes(:,[54,55,58,59]) = repmat([f;f;f],[1,4]); % Fix z

prenodes(:,[32,48]) = repmat([0;0;e],[1,2]); % Strain x, fix y

prenodes(:,[56,60]) = repmat([f;0;e],[1,2]); % Fix y,z

prenodes(:,[62,63]) = repmat([0;f;f],[1,2]); % Strain x, fix z

Page 385: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

356

% Calculated Corners

prenodes(:,16) = prenodes(:,13) + prenodes(:, 4) - prenodes(:, 1);

prenodes(:,52) = prenodes(:,49) + prenodes(:, 4) - prenodes(:, 1);

prenodes(:,61) = prenodes(:,49) + prenodes(:,13) - prenodes(:, 1);

prenodes(:,64) = prenodes(:,49) + prenodes(:,13) + prenodes(:, 4) -

2*prenodes(:,1);

% Generate PRDOF showing strains and index for removing prescribed dofs

for i = 1:size(prenodes,2)

for j = 1:3

dof = (3*(i-1) + j);

bc = prenodes(j,i);

if (bc == 0 || bc == e)

INDEX(4,dof) = false;

PRDOF(4,dof) = bc;

end

end

end

%Prepare Force Vector

F4 = -K1_4*(PRDOF(4,:)');

% for i = 1:length(a1)

% F4(b1(i)) = F4(b1(i)) + F4(a1(i));

% end

% Reduce vectors and matrices based on prescribed and periodic

nn4 = nn(INDEX(4,:));

F4_r = F4(INDEX(4,:)); % Reduced Force vector

K1_4r = K1_4(INDEX(4,:),:);

K1_4r = K1_4r(:,INDEX(4,:));

%% Test 5

% Prescribed BC's

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,49) = [e;0;0]; % Strain in x, else fix

prenodes(:,[1,4,13]) = repmat([0;0;0],[1,3]); % All fixed

prenodes(:,[24,28,40,44]) = repmat([f;0;f],[1,4]); % Fix y

prenodes(:,[30,31,46,47]) = repmat([f;f;f],[1,4]); % Strain x

prenodes(:,[54,55,58,59]) = repmat([e;f;0],[1,4]); % Fix z

prenodes(:,[32,48]) = repmat([f;0;f],[1,2]); % Strain x, fix y

prenodes(:,[56,60]) = repmat([e;0;0],[1,2]); % Fix y,z

prenodes(:,[62,63]) = repmat([e;f;0],[1,2]); % Strain x, fix z

% Calculated Corners

prenodes(:,16) = prenodes(:,13) + prenodes(:, 4) - prenodes(:, 1);

prenodes(:,52) = prenodes(:,49) + prenodes(:, 4) - prenodes(:, 1);

prenodes(:,61) = prenodes(:,49) + prenodes(:,13) - prenodes(:, 1);

prenodes(:,64) = prenodes(:,49) + prenodes(:,13) + prenodes(:, 4) -

2*prenodes(:,1);

% Generate PRDOF showing strains and index for removing prescribed dofs

for i = 1:size(prenodes,2)

for j = 1:3

dof = (3*(i-1) + j);

bc = prenodes(j,i);

if (bc == 0 || bc == e)

INDEX(5,dof) = false;

PRDOF(5,dof) = bc;

end

end

Page 386: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

357

end

%Prepare Force Vector

F5 = -K1_5*(PRDOF(5,:)');

% for i = 1:length(a1)

% F5(b1(i)) = F5(b1(i)) + F5(a1(i));

% end

% Reduce vectors and matrices based on prescribed and periodic

nn5 = nn(INDEX(5,:));

F5_r = F5(INDEX(5,:)); % Reduced Force vector

K1_5r = K1_5(INDEX(5,:),:);

K1_5r = K1_5r(:,INDEX(5,:));

%% Test 6

% Prescribed BC's

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,13) = [0;e;0]; % Strain in x, else fix

prenodes(:,[1,4,49]) = repmat([0;0;0],[1,3]); % All fixed

prenodes(:,[24,28,40,44]) = repmat([f;f;f],[1,4]); % Fix y

prenodes(:,[30,31,46,47]) = repmat([0;e;f],[1,4]); % Strain x

prenodes(:,[54,55,58,59]) = repmat([f;f;0],[1,4]); % Fix z

prenodes(:,[32,48]) = repmat([0;e;f],[1,2]); % Strain x, fix y

prenodes(:,[56,60]) = repmat([f;f;0],[1,2]); % Fix y,z

prenodes(:,[62,63]) = repmat([0;e;0],[1,2]); % Strain x, fix z

% Calculated Corners

prenodes(:,16) = prenodes(:,13) + prenodes(:, 4) - prenodes(:, 1);

prenodes(:,52) = prenodes(:,49) + prenodes(:, 4) - prenodes(:, 1);

prenodes(:,61) = prenodes(:,49) + prenodes(:,13) - prenodes(:, 1);

prenodes(:,64) = prenodes(:,49) + prenodes(:,13) + prenodes(:, 4) -

2*prenodes(:,1);

% Generate PRDOF showing strains and index for removing prescribed dofs

for i = 1:size(prenodes,2)

for j = 1:3

dof = (3*(i-1) + j);

bc = prenodes(j,i);

if (bc == 0 || bc == e)

INDEX(6,dof) = false;

PRDOF(6,dof) = bc;

end

end

end

%Prepare Force Vector

F6 = -K1_6*(PRDOF(6,:)');

% for i = 1:length(a1)

% F6(b1(i)) = F6(b1(i)) + F6(a1(i));

% end

% Reduce vectors and matrices based on prescribed and periodic

nn6 = nn(INDEX(6,:));

F6_r = F6(INDEX(6,:)); % Reduced Force vector

K1_6r = K1_6(INDEX(6,:),:);

K1_6r = K1_6r(:,INDEX(6,:));

%% Solve all test cases and rebuild displacement vectors

% Solve FEA

ua1_r = K1_1r\F1_r;

Page 387: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

358

ua2_r = K1_2r\F2_r;

ua3_r = K1_3r\F3_r;

ua4_r = K1_4r\F4_r;

ua5_r = K1_5r\F5_r;

ua6_r = K1_6r\F6_r;

% Fill in prescribed/calculated dof

UA = [PRDOF(1,:);

PRDOF(2,:);

PRDOF(3,:);

PRDOF(4,:);

PRDOF(5,:);

PRDOF(6,:)];

% Fill in solved dof

UA(1,nn1) = ua1_r;

UA(2,nn2) = ua2_r;

UA(3,nn3) = ua3_r;

UA(4,nn4) = ua4_r;

UA(5,nn5) = ua5_r;

UA(6,nn6) = ua6_r;

% Calculate and fill in periodic displacements

for i = 1:length(a1)

UA(:,a1(i)) = UA(:,b1(i)) - UA(:,c1(i)) + UA(:,d1(i));

end

% Reshape to xy solutions

UA_XY = [reshape(UA(1,:),[3,gs.Ndof/3]);

reshape(UA(2,:),[3,gs.Ndof/3]);

reshape(UA(3,:),[3,gs.Ndof/3]);

reshape(UA(4,:),[3,gs.Ndof/3]);

reshape(UA(5,:),[3,gs.Ndof/3]);

reshape(UA(6,:),[3,gs.Ndof/3])];

%% Calculate Element Mutual Energies

D = zeros(36,gs.M);

% Prepare induced displacement matrices

for i = 1:gs.M

nodeA = gs.IEN(1,i);

nodeB = gs.IEN(2,i);

D( 1:6 ,i) = [UA_XY(1:3,nodeA); UA_XY(1:3,nodeB)];

D( 7:12,i) = [UA_XY(4:6,nodeA); UA_XY(4:6,nodeB)];

D(13:18,i) = [UA_XY(7:9,nodeA); UA_XY(7:9,nodeB)];

D(19:24,i) = [UA_XY(10:12,nodeA); UA_XY(10:12,nodeB)];

D(25:30,i) = [UA_XY(13:15,nodeA); UA_XY(13:15,nodeB)];

D(31:36,i) = [UA_XY(16:18,nodeA); UA_XY(16:18,nodeB)];

end

%%%% Compute Elasticity Matrix

A_min = min(A);

A_rel = find(A>A_min);

Y = opt.ucs(1);

q = zeros(6);

for h = 1:gs.M

for i = 1:6

for j = 1:6

val = (1/Y)*(D((6*i-5):(6*i),h))'*...

k_all(:,(6*h-5):(6*h))*(D((6*j-5):(6*j),h));

Page 388: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

359

q(i,j) = q(i,j) + val;

end

end

end

Eh = q;

%%% Calculate Buckling Load

Kbuck = 1;

F_crit = -

pi*pi*opt.mod*(Abuck(A_rel)'.*Abuck(A_rel)'/4/pi)./(Kbuck*gs.L(A_rel)).^2;

F_crit = gs.surfred(A_rel)'.*F_crit; % Account for surface

L_chng = (sum((D(16:18,A_rel)-D(13:15,A_rel)).^2).^(.5))'; % Deformed length

F_axi = Abuck(A_rel)'*opt.mod.*(L_chng)./gs.L(A_rel); % Axial force in each

member

factor2crit = -F_crit./F_axi;

F_eff = (opt.ucs(1)^2)*Eh(3,3)*(e)/opt.ucs(1);

min2buckle = F_eff*min(factor2crit(factor2crit>0));

% ALTO3D_plotdeformed(gs.Ainitial,gs,UA_XY(1:3,:)+gs.coord,A_relevant)

% ALTO3D_plotdeformed(gs.Ainitial,gs,UA_XY(4:6,:)+gs.coord,A_relevant)

% ALTO3D_plotdeformed(gs.Ainitial,gs,UA_XY(7:9,:)+gs.coord,A_relevant)

% ALTO3D_plotdeformed(gs.Ainitial,gs,UA_XY(10:12,:)+gs.coord,A_relevant)

% ALTO3D_plotdeformed(gs.Ainitial,gs,UA_XY(13:15,:)+gs.coord,A_relevant)

% ALTO3D_plotdeformed(gs.Ainitial,gs,UA_XY(16:18,:)+gs.coord,A_relevant)

end

ALTO3D_homog5.m

function [Eh,min2buckle] = ALTO3D_homog5(A,Abuck,opt,optim,gs)

A_relevant = (pi/4*(opt.drel)^2);

A_rel = A>=A_relevant;

% local member stiffness = AE/L

% element stiffness matrix = k

% global stiffness matrix for part 1 = K1

K1=zeros(gs.Ndof,gs.Ndof);

k_all=zeros(6,6*gs.M);

dk1v=zeros(36,gs.M);

for i=1:gs.M

Cx = gs.theta(i,1);

Cy = gs.theta(i,2);

Cz = gs.theta(i,3);

lambda = [ Cx*Cx Cx*Cy Cx*Cz;

Cx*Cy Cy*Cy Cy*Cz;

Cx*Cz Cy*Cz Cz*Cz];

k = [ lambda, -lambda;

-lambda, lambda];

dk = opt.mod/gs.L(i)*k;

k = A(i)*opt.mod/gs.L(i)*k;

dk1v(:,i) = reshape(dk',[36,1]);

k_all(1:6,(6*i-5):(6*i)) = k;

q = gs.LM(:,i);

Page 389: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

360

K1(q,q) = K1(q,q) + k;

end

%% Modify stiffness matrix for periodic BC's for each strain test

%%% Unit test strain cases 1 and 2 (axial strain), 3 (shear strain)

K1_1 = K1;

%%%% Boundary Conditons %%%%

% This code is for reducing K1 using the following equations where the

% numbers are the corresponding degree of freedom. These equations come

% from the following general periodicity equations (van der Sluis 2000):

% u_11 = u_12 - u_v4

% u_22 = u_21 - u_v1

% u_v3 = u_v2 + u_v4 - u_v1

%

% Using the general form

% a = b - c + d -or- a = b - c + d + e1 - e2

% Symmetry face nodes % CAREFUL WHICH NODES ARE CALCULATED FIRST, 14/15

% must be calculated before 2/3 can be calculated

xnodes = [27,28,29,30,52,53,54,55,77,78,79, 80,102,103,104;

47,48,49,50,72,73,74,75,97,98,99,100,122,123,124;

21,21,21,21,21,21,21,21,21,21,21, 21, 21, 21, 21;

1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1];

ynodes = [31,36,41,46,56,61,66,71,81,86,91, 96,106,111,116;

35,40,45,50,60,65,70,75,85,90,95,100,110,115,120;

5, 5, 5, 5, 5, 5, 5, 5, 5, 5, 5, 5, 5, 5, 5;

1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1];

znodes = [ 7, 8, 9, 10, 12, 13, 14, 15, 17, 18, 19, 20, 22, 23, 24;

107,108,109,110,112,113,114,115,117,118,119,120,122,123,124;

101,101,101,101,101,101,101,101,101,101,101,101,101,101,101;

1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1, 1];

others = [ 2, 3, 4, 6, 11, 16,26,51,76;

22,23,24,106,111,116,30,55,80;

21,21,21,101,101,101, 5, 5, 5;

1, 1, 1, 1, 1, 1, 1, 1, 1];

nodes = [ xnodes, ynodes, znodes, others];

a1 = zeros(1,54);

b1 = zeros(1,54);

c1 = zeros(1,54);

d1 = zeros(1,54);

for i=1:length(nodes(1,:))

a1((3*i-2):3*i) = (3*nodes(1,i)-2):3*nodes(1,i);

b1((3*i-2):3*i) = (3*nodes(2,i)-2):3*nodes(2,i);

c1((3*i-2):3*i) = (3*nodes(3,i)-2):3*nodes(3,i);

d1((3*i-2):3*i) = (3*nodes(4,i)-2):3*nodes(4,i);

end

for i = 1:length(a1)

K1_1(:,b1(i)) = K1_1(:,b1(i)) + K1_1(:,a1(i));

K1_1(:,c1(i)) = K1_1(:,c1(i)) - K1_1(:,a1(i));

Page 390: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

361

K1_1(:,d1(i)) = K1_1(:,d1(i)) + K1_1(:,a1(i));

end

% for i = 1:length(a1)

% K1_1(b1(i),:) = K1_1(b1(i),:) + K1_1(a1(i),:); % MISTAKE? %

% end

K1_2 = K1_1;

K1_3 = K1_1;

K1_4 = K1_1;

K1_5 = K1_1;

K1_6 = K1_1;

% % Prescribed Displacements

pd1 = zeros(gs.Ndof,1);

pd2 = zeros(gs.Ndof,1);

pd3 = zeros(gs.Ndof,1);

pd4 = zeros(gs.Ndof,1);

pd5 = zeros(gs.Ndof,1);

pd6 = zeros(gs.Ndof,1);

nn = (1:gs.Ndof)';

INDEX = true(6,gs.Ndof); % All dof that will be removed

INDEX(:,a1) = false; % Set periodic dofs to be removed

PRDOF = zeros(6,gs.Ndof); % All dof with strain

% Strain (e); Free (f); Fixed (0);

e = 1;

f = 2;

bctype = 1; % 1 is original, 2 is modified but not periodic corners

%% Test 1

% Prescribed BC's

if bctype == 1

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,21) = [e;0;0]; % Strain in x, else fix

prenodes(:,[1,5,101]) = repmat([0;0;0],[1,3]); % All fixed

prenodes(:,[ 35, 40, 45, 60, 65, 70, 85, 90, 95]) = repmat([f;0;f],[1,9]); %

Fix y

prenodes(:,[ 47, 48, 49, 72, 73, 74, 97, 98, 99]) = repmat([e;f;f],[1,9]); %

Strain x

prenodes(:,[107,108,109,112,113,114,117,118,119]) = repmat([f;f;0],[1,9]); %

Fix z

prenodes(:,[ 50, 75,100]) = repmat([e;0;f],[1,3]); % Strain x, fix y

prenodes(:,[110,115,120]) = repmat([f;0;0],[1,3]); % Fix y,z

prenodes(:,[122,123,124]) = repmat([e;f;0],[1,3]); % Strain x, fix z

else

% Prescribed BC's NEW

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,21) = [e;0;0]; % Strain in x, else fix

prenodes(:,1) = [0;0;0]; % All fixed

prenodes(:,5) = [0;f;0]; % All fixed

prenodes(:,101) = [0;0;f]; % All fixed

prenodes(:,[ 35, 40, 45, 60, 65, 70, 85, 90, 95]) = repmat([f;f;f],[1,9]); %

Fix y

prenodes(:,[ 47, 48, 49, 72, 73, 74, 97, 98, 99]) = repmat([e;f;f],[1,9]); %

Strain x

prenodes(:,[107,108,109,112,113,114,117,118,119]) = repmat([f;f;f],[1,9]); %

Fix z

prenodes(:,[ 50, 75,100]) = repmat([e;f;f],[1,3]); % Strain x, fix y

prenodes(:,[110,115,120]) = repmat([f;f;f],[1,3]); % Fix y,z

prenodes(:,[122,123,124]) = repmat([e;f;f],[1,3]); % Strain x, fix z

Page 391: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

362

end

% Calculated Corners

prenodes(:,25) = prenodes(:, 21) + prenodes(:, 5) - prenodes(:, 1);

prenodes(:,105) = prenodes(:,101) + prenodes(:, 5) - prenodes(:, 1);

prenodes(:,121) = prenodes(:,101) + prenodes(:,21) - prenodes(:, 1);

prenodes(:,125) = prenodes(:,101) + prenodes(:,21) + prenodes(:, 5) -

2*prenodes(:,1);

% Generate PRDOF showing strains and index for removing prescribed dofs

for i = 1:size(prenodes,2)

for j = 1:3

dof = (3*(i-1) + j);

bc = prenodes(j,i);

if (bc == 0 || bc == e)

INDEX(1,dof) = false;

PRDOF(1,dof) = bc;

end

end

end

%Prepare Force Vector

F1 = -K1_1*(PRDOF(1,:)');

% for i = 1:length(a1) % MISTAKE? %

% F1(b1(i)) = F1(b1(i)) + F1(a1(i));

% end

% Reduce vectors and matrices based on prescribed and periodic

nn1 = nn(INDEX(1,:));

F1_r = F1(INDEX(1,:)); % Reduced Force vector

K1_1r = K1_1(INDEX(1,:),:);

K1_1r = K1_1r(:,INDEX(1,:));

%% Test 2

if bctype ==1

% Prescribed BC's

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,5) = [0;e;0]; % Strain in y, else fix

prenodes(:,[1,21,101]) = repmat([0;0;0],[1,3]); % Fix all

prenodes(:,[ 35, 40, 45, 60, 65, 70, 85, 90, 95]) = repmat([f;e;f],[1,9]); %

Free x,z Strain y

prenodes(:,[ 47, 48, 49, 72, 73, 74, 97, 98, 99]) = repmat([0;f;f],[1,9]); %

Fix x Free y,z

prenodes(:,[107,108,109,112,113,114,117,118,119]) = repmat([f;f;0],[1,9]); %

Free x,y Fix z

prenodes(:,[ 50, 75,100]) = repmat([0;e;f],[1,3]); % Fix x Strain y Free

z

prenodes(:,[110,115,120]) = repmat([f;e;0],[1,3]); % Free x Strain y Fix

z

prenodes(:,[122,123,124]) = repmat([0;f;0],[1,3]); % Fix x Free y Fix z

else

% Prescribed BC's

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,5) = [0;e;0]; % Strain in y, else fix

prenodes(:,1) = [0;0;0]; % All fixed

prenodes(:,21) = [f;0;0]; % All fixed

prenodes(:,101) = [0;0;f]; % All fixed

prenodes(:,[ 35, 40, 45, 60, 65, 70, 85, 90, 95]) = repmat([f;e;f],[1,9]); %

Free x,z Strain y

prenodes(:,[ 47, 48, 49, 72, 73, 74, 97, 98, 99]) = repmat([f;f;f],[1,9]); %

Fix x Free y,z

Page 392: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

363

prenodes(:,[107,108,109,112,113,114,117,118,119]) = repmat([f;f;f],[1,9]); %

Free x,y Fix z

prenodes(:,[ 50, 75,100]) = repmat([f;e;f],[1,3]); % Fix x Strain y Free

z

prenodes(:,[110,115,120]) = repmat([f;e;f],[1,3]); % Free x Strain y Fix

z

prenodes(:,[122,123,124]) = repmat([f;f;f],[1,3]); % Fix x Free y Fix z

end

% Calculated Corners

prenodes(:,25) = prenodes(:, 21) + prenodes(:, 5) - prenodes(:, 1);

prenodes(:,105) = prenodes(:,101) + prenodes(:, 5) - prenodes(:, 1);

prenodes(:,121) = prenodes(:,101) + prenodes(:,21) - prenodes(:, 1);

prenodes(:,125) = prenodes(:,101) + prenodes(:,21) + prenodes(:, 5) -

2*prenodes(:,1);

% Generate PRDOF showing strains and index for removing prescribed dofs

for i = 1:size(prenodes,2)

for j = 1:3

dof = (3*(i-1) + j);

bc = prenodes(j,i);

if (bc == 0 || bc == e)

INDEX(2,dof) = false;

PRDOF(2,dof) = bc;

end

end

end

%Prepare Force Vector

F2 = -K1_2*(PRDOF(2,:)');

% for i = 1:length(a1) % MISTAKE? %

% F2(b1(i)) = F2(b1(i)) + F2(a1(i));

% end

% Reduce vectors and matrices based on prescribed and periodic

nn2 = nn(INDEX(2,:));

F2_r = F2(INDEX(2,:)); % Reduced Force vector

K1_2r = K1_2(INDEX(2,:),:);

K1_2r = K1_2r(:,INDEX(2,:));

%% Test 3

if bctype ==1

% Prescribed BC's

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,101) = [0;0;e]; % Strain in x, else fix

prenodes(:,[1,5,21]) = repmat([0;0;0],[1,3]); % All fixed

prenodes(:,[ 35, 40, 45, 60, 65, 70, 85, 90, 95]) = repmat([f;0;f],[1,9]); %

Fix y

prenodes(:,[ 47, 48, 49, 72, 73, 74, 97, 98, 99]) = repmat([0;f;f],[1,9]); %

Strain x

prenodes(:,[107,108,109,112,113,114,117,118,119]) = repmat([f;f;e],[1,9]); %

Fix z

prenodes(:,[ 50, 75,100]) = repmat([0;0;f],[1,3]); % Strain x, fix y

prenodes(:,[110,115,120]) = repmat([f;0;e],[1,3]); % Fix y,z

prenodes(:,[122,123,124]) = repmat([0;f;e],[1,3]); % Strain x, fix z

else

% Prescribed BC's

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,101) = [0;0;e]; % Strain in x, else fix

prenodes(:,1) = [0;0;0]; % All fixed

prenodes(:,5) = [0;f;0]; % All fixed

Page 393: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

364

prenodes(:,21) = [f;0;0]; % All fixed

prenodes(:,[ 35, 40, 45, 60, 65, 70, 85, 90, 95]) = repmat([f;f;f],[1,9]); %

Fix y

prenodes(:,[ 47, 48, 49, 72, 73, 74, 97, 98, 99]) = repmat([f;f;f],[1,9]); %

Strain x

prenodes(:,[107,108,109,112,113,114,117,118,119]) = repmat([f;f;e],[1,9]); %

Fix z

prenodes(:,[ 50, 75,100]) = repmat([f;f;f],[1,3]); % Strain x, fix y

prenodes(:,[110,115,120]) = repmat([f;f;e],[1,3]); % Fix y,z

prenodes(:,[122,123,124]) = repmat([f;f;e],[1,3]); % Strain x, fix z

end

% Calculated Corners

prenodes(:,25) = prenodes(:, 21) + prenodes(:, 5) - prenodes(:, 1);

prenodes(:,105) = prenodes(:,101) + prenodes(:, 5) - prenodes(:, 1);

prenodes(:,121) = prenodes(:,101) + prenodes(:,21) - prenodes(:, 1);

prenodes(:,125) = prenodes(:,101) + prenodes(:,21) + prenodes(:, 5) -

2*prenodes(:,1);

% Generate PRDOF showing strains and index for removing prescribed dofs

for i = 1:size(prenodes,2)

for j = 1:3

dof = (3*(i-1) + j);

bc = prenodes(j,i);

if (bc == 0 || bc == e)

INDEX(3,dof) = false;

PRDOF(3,dof) = bc;

end

end

end

%Prepare Force Vector

F3 = -K1_3*(PRDOF(3,:)');

% for i = 1:length(a1) % MISTAKE? %

% F3(b1(i)) = F3(b1(i)) + F3(a1(i));

% end

% Reduce vectors and matrices based on prescribed and periodic

nn3 = nn(INDEX(3,:));

F3_r = F3(INDEX(3,:)); % Reduced Force vector

K1_3r = K1_3(INDEX(3,:),:);

K1_3r = K1_3r(:,INDEX(3,:));

%% Test 4

if bctype == 1

% Prescribed BC's

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,5) = [0;0;e]; % Strain in x, else fix

prenodes(:,[1,21,101]) = repmat([0;0;0],[1,3]); % All fixed

prenodes(:,[ 35, 40, 45, 60, 65, 70, 85, 90, 95]) = repmat([f;0;e],[1,9]); %

Fix y

prenodes(:,[ 47, 48, 49, 72, 73, 74, 97, 98, 99]) = repmat([0;f;f],[1,9]); %

Strain x

prenodes(:,[107,108,109,112,113,114,117,118,119]) = repmat([f;f;f],[1,9]); %

Fix z

prenodes(:,[ 50, 75,100]) = repmat([0;0;e],[1,3]); % Strain x, fix y

prenodes(:,[110,115,120]) = repmat([f;0;e],[1,3]); % Fix y,z

prenodes(:,[122,123,124]) = repmat([0;f;f],[1,3]); % Strain x, fix z

else

% Prescribed BC's

prenodes = f*ones(3,gs.N); % Start with all dof free

Page 394: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

365

prenodes(:,5) = [0;0;e]; % Strain in x, else fix

prenodes(:,[1,21,101]) = repmat([0;0;0],[1,3]); % All fixed

prenodes(:,[ 35, 40, 45, 60, 65, 70, 85, 90, 95]) = repmat([f;f;e],[1,9]); %

Fix y

prenodes(:,[ 47, 48, 49, 72, 73, 74, 97, 98, 99]) = repmat([f;f;f],[1,9]); %

Strain x

prenodes(:,[107,108,109,112,113,114,117,118,119]) = repmat([f;f;f],[1,9]); %

Fix z

prenodes(:,[ 50, 75,100]) = repmat([f;f;e],[1,3]); % Strain x, fix y

prenodes(:,[110,115,120]) = repmat([f;f;e],[1,3]); % Fix y,z

prenodes(:,[122,123,124]) = repmat([f;f;f],[1,3]); % Strain x, fix z

end

% Calculated Corners

prenodes(:,25) = prenodes(:, 21) + prenodes(:, 5) - prenodes(:, 1);

prenodes(:,105) = prenodes(:,101) + prenodes(:, 5) - prenodes(:, 1);

prenodes(:,121) = prenodes(:,101) + prenodes(:,21) - prenodes(:, 1);

prenodes(:,125) = prenodes(:,101) + prenodes(:,21) + prenodes(:, 5) -

2*prenodes(:,1);

% Generate PRDOF showing strains and index for removing prescribed dofs

for i = 1:size(prenodes,2)

for j = 1:3

dof = (3*(i-1) + j);

bc = prenodes(j,i);

if (bc == 0 || bc == e)

INDEX(4,dof) = false;

PRDOF(4,dof) = bc;

end

end

end

%Prepare Force Vector

F4 = -K1_4*(PRDOF(4,:)');

% for i = 1:length(a1) % MISTAKE? %

% F4(b1(i)) = F4(b1(i)) + F4(a1(i));

% end

% Reduce vectors and matrices based on prescribed and periodic

nn4 = nn(INDEX(4,:));

F4_r = F4(INDEX(4,:)); % Reduced Force vector

K1_4r = K1_4(INDEX(4,:),:);

K1_4r = K1_4r(:,INDEX(4,:));

%% Test 5

if bctype == 1

% Prescribed BC's

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,101) = [e;0;0]; % Strain in x, else fix

prenodes(:,[1,5,21]) = repmat([0;0;0],[1,3]); % All fixed

prenodes(:,[ 35, 40, 45, 60, 65, 70, 85, 90, 95]) = repmat([f;0;f],[1,9]); %

Fix y

prenodes(:,[ 47, 48, 49, 72, 73, 74, 97, 98, 99]) = repmat([f;f;f],[1,9]); %

Strain x

prenodes(:,[107,108,109,112,113,114,117,118,119]) = repmat([e;f;0],[1,9]); %

Fix z

prenodes(:,[ 50, 75,100]) = repmat([f;0;f],[1,3]); % Strain x, fix y

prenodes(:,[110,115,120]) = repmat([e;0;0],[1,3]); % Fix y,z

prenodes(:,[122,123,124]) = repmat([e;f;0],[1,3]); % Strain x, fix z

else

% Prescribed BC's

Page 395: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

366

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,101) = [e;0;0]; % Strain in x, else fix

prenodes(:,[1,5,21]) = repmat([0;0;0],[1,3]); % All fixed

prenodes(:,[ 35, 40, 45, 60, 65, 70, 85, 90, 95]) = repmat([f;f;f],[1,9]); %

Fix y

prenodes(:,[ 47, 48, 49, 72, 73, 74, 97, 98, 99]) = repmat([f;f;f],[1,9]); %

Strain x

prenodes(:,[107,108,109,112,113,114,117,118,119]) = repmat([e;f;f],[1,9]); %

Fix z

prenodes(:,[ 50, 75,100]) = repmat([f;f;f],[1,3]); % Strain x, fix y

prenodes(:,[110,115,120]) = repmat([e;f;f],[1,3]); % Fix y,z

prenodes(:,[122,123,124]) = repmat([e;f;f],[1,3]); % Strain x, fix z

end

% Calculated Corners

prenodes(:,25) = prenodes(:, 21) + prenodes(:, 5) - prenodes(:, 1);

prenodes(:,105) = prenodes(:,101) + prenodes(:, 5) - prenodes(:, 1);

prenodes(:,121) = prenodes(:,101) + prenodes(:,21) - prenodes(:, 1);

prenodes(:,125) = prenodes(:,101) + prenodes(:,21) + prenodes(:, 5) -

2*prenodes(:,1);

% Generate PRDOF showing strains and index for removing prescribed dofs

for i = 1:size(prenodes,2)

for j = 1:3

dof = (3*(i-1) + j);

bc = prenodes(j,i);

if (bc == 0 || bc == e)

INDEX(5,dof) = false;

PRDOF(5,dof) = bc;

end

end

end

%Prepare Force Vector

F5 = -K1_5*(PRDOF(5,:)');

% for i = 1:length(a1) % MISTAKE? %

% F5(b1(i)) = F5(b1(i)) + F5(a1(i));

% end

% Reduce vectors and matrices based on prescribed and periodic

nn5 = nn(INDEX(5,:));

F5_r = F5(INDEX(5,:)); % Reduced Force vector

K1_5r = K1_5(INDEX(5,:),:);

K1_5r = K1_5r(:,INDEX(5,:));

%% Test 6

if bctype == 1

% Prescribed BC's

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,21) = [0;e;0]; % Strain in x, else fix

prenodes(:,[1,5,101]) = repmat([0;0;0],[1,3]); % All fixed

prenodes(:,[ 35, 40, 45, 60, 65, 70, 85, 90, 95]) = repmat([f;f;f],[1,9]); %

Fix y

prenodes(:,[ 47, 48, 49, 72, 73, 74, 97, 98, 99]) = repmat([0;e;f],[1,9]); %

Strain x

prenodes(:,[107,108,109,112,113,114,117,118,119]) = repmat([f;f;0],[1,9]); %

Fix z

prenodes(:,[ 50, 75,100]) = repmat([0;e;f],[1,3]); % Strain x, fix y

prenodes(:,[110,115,120]) = repmat([f;f;0],[1,3]); % Fix y,z

prenodes(:,[122,123,124]) = repmat([0;e;0],[1,3]); % Strain x, fix z

else

Page 396: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

367

% Prescribed BC's

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,21) = [0;e;0]; % Strain in x, else fix

prenodes(:,[1,5,101]) = repmat([0;0;0],[1,3]); % All fixed

prenodes(:,[ 35, 40, 45, 60, 65, 70, 85, 90, 95]) = repmat([f;f;f],[1,9]); %

Fix y

prenodes(:,[ 47, 48, 49, 72, 73, 74, 97, 98, 99]) = repmat([f;e;f],[1,9]); %

Strain x

prenodes(:,[107,108,109,112,113,114,117,118,119]) = repmat([f;f;f],[1,9]); %

Fix z

prenodes(:,[ 50, 75,100]) = repmat([f;e;f],[1,3]); % Strain x, fix y

prenodes(:,[110,115,120]) = repmat([f;f;f],[1,3]); % Fix y,z

prenodes(:,[122,123,124]) = repmat([f;e;f],[1,3]); % Strain x, fix z

end

% Calculated Corners

prenodes(:,25) = prenodes(:, 21) + prenodes(:, 5) - prenodes(:, 1);

prenodes(:,105) = prenodes(:,101) + prenodes(:, 5) - prenodes(:, 1);

prenodes(:,121) = prenodes(:,101) + prenodes(:,21) - prenodes(:, 1);

prenodes(:,125) = prenodes(:,101) + prenodes(:,21) + prenodes(:, 5) -

2*prenodes(:,1);

% Generate PRDOF showing strains and index for removing prescribed dofs

for i = 1:size(prenodes,2)

for j = 1:3

dof = (3*(i-1) + j);

bc = prenodes(j,i);

if (bc == 0 || bc == e)

INDEX(6,dof) = false;

PRDOF(6,dof) = bc;

end

end

end

%Prepare Force Vector

F6 = -K1_6*(PRDOF(6,:)');

% for i = 1:length(a1) % MISTAKE? %

% F6(b1(i)) = F6(b1(i)) + F6(a1(i));

% end

% Reduce vectors and matrices based on prescribed and periodic

nn6 = nn(INDEX(6,:));

F6_r = F6(INDEX(6,:)); % Reduced Force vector

K1_6r = K1_6(INDEX(6,:),:);

K1_6r = K1_6r(:,INDEX(6,:));

%% Solve all test cases and rebuild displacement vectors

% Solve FEA

ua1_r = K1_1r\F1_r;

ua2_r = K1_2r\F2_r;

ua3_r = K1_3r\F3_r;

ua4_r = K1_4r\F4_r;

ua5_r = K1_5r\F5_r;

ua6_r = K1_6r\F6_r;

% Fill in prescribed/calculated dof

UA = [PRDOF(1,:);

PRDOF(2,:);

PRDOF(3,:);

PRDOF(4,:);

Page 397: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

368

PRDOF(5,:);

PRDOF(6,:)];

% Fill in solved dof

UA(1,nn1) = ua1_r;

UA(2,nn2) = ua2_r;

UA(3,nn3) = ua3_r;

UA(4,nn4) = ua4_r;

UA(5,nn5) = ua5_r;

UA(6,nn6) = ua6_r;

% Calculate and fill in periodic displacements

for i = 1:length(a1)

UA(:,a1(i)) = UA(:,b1(i)) - UA(:,c1(i)) + UA(:,d1(i));

end

% Reshape to xy solutions

UA_XY = [reshape(UA(1,:),[3,gs.Ndof/3]);

reshape(UA(2,:),[3,gs.Ndof/3]);

reshape(UA(3,:),[3,gs.Ndof/3]);

reshape(UA(4,:),[3,gs.Ndof/3]);

reshape(UA(5,:),[3,gs.Ndof/3]);

reshape(UA(6,:),[3,gs.Ndof/3])];

% Calculate new lengths

newcoord = repmat(gs.coord,6,1) + UA_XY;

for i = 1:gs.M

nodeA = gs.IEN(1,i);

nodeB = gs.IEN(2,i);

pos1 = newcoord(1:3,nodeA);

pos2 = newcoord(1:3,nodeB);

gs.Ldef(i,:) = sqrt(sum((pos2-pos1).^2));

end

%% Calculate Element Mutual Energies

D = zeros(36,gs.M);

% Prepare induced displacement matrices

for i = 1:gs.M

nodeA = gs.IEN(1,i);

nodeB = gs.IEN(2,i);

D( 1:6 ,i) = [UA_XY(1:3,nodeA); UA_XY(1:3,nodeB)];

D( 7:12,i) = [UA_XY(4:6,nodeA); UA_XY(4:6,nodeB)];

D(13:18,i) = [UA_XY(7:9,nodeA); UA_XY(7:9,nodeB)];

D(19:24,i) = [UA_XY(10:12,nodeA); UA_XY(10:12,nodeB)];

D(25:30,i) = [UA_XY(13:15,nodeA); UA_XY(13:15,nodeB)];

D(31:36,i) = [UA_XY(16:18,nodeA); UA_XY(16:18,nodeB)];

end

%%%% Compute Elasticity Matrix

A_min = min(A);

A_rel = find(A>A_min);

Y = opt.ucs(1);

q = zeros(6);

for h = 1:gs.M

for i = 1:6

for j = 1:6

val = (1/Y)*(D((6*i-5):(6*i),h))'*...

Page 398: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

369

k_all(:,(6*h-5):(6*h))*(D((6*j-5):(6*j),h));

q(i,j) = q(i,j) + val;

end

end

end

Eh = q;

%%% Calculate Buckling Load

Kbuck = 1;

F_crit = -

pi*pi*opt.mod*(Abuck(A_rel)'.*Abuck(A_rel)'/4/pi)./(Kbuck*gs.L(A_rel)).^2;

F_crit = gs.surfred(A_rel)'.*F_crit; % Account for surface

L_chng = (sum((D(16:18,A_rel)-D(13:15,A_rel)).^2).^(.5))'; % Deformed length

F_axi = Abuck(A_rel)'*opt.mod.*(L_chng)./gs.L(A_rel); % Axial force in each

member

factor2crit = -F_crit./F_axi;

F_eff = (opt.ucs(1)^2)*Eh(3,3)*(e)/opt.ucs(1);

min2buckle = F_eff*min(factor2crit(factor2crit>0));

% ALTO3D_plotdeformed(gs.Ainitial,gs,UA_XY(1:3,:)+gs.coord,A_relevant)

% ALTO3D_plotdeformed(gs.Ainitial,gs,UA_XY(4:6,:)+gs.coord,A_relevant)

% ALTO3D_plotdeformed(gs.Ainitial,gs,UA_XY(7:9,:)+gs.coord,A_relevant)

% ALTO3D_plotdeformed(gs.Ainitial,gs,UA_XY(10:12,:)+gs.coord,A_relevant)

% ALTO3D_plotdeformed(gs.Ainitial,gs,UA_XY(13:15,:)+gs.coord,A_relevant)

% ALTO3D_plotdeformed(gs.Ainitial,gs,UA_XY(16:18,:)+gs.coord,A_relevant)

end

ALTO3D_homog5v2.m

function [Eh,min2buckle] = ALTO3D_homog5v2(A,Abuck,opt,optim,gs)

A_relevant = (pi/4*(opt.drel)^2);

A_rel = A>=A_relevant;

% local member stiffness = AE/L

% element stiffness matrix = k

% global stiffness matrix for part 1 = K1

K1=zeros(gs.Ndof,gs.Ndof);

k_all=zeros(6,6*gs.M);

dk1v=zeros(36,gs.M);

for i=1:gs.M

Cx = gs.theta(i,1);

Cy = gs.theta(i,2);

Cz = gs.theta(i,3);

lambda = [ Cx*Cx Cx*Cy Cx*Cz;

Cx*Cy Cy*Cy Cy*Cz;

Cx*Cz Cy*Cz Cz*Cz];

k = [ lambda, -lambda;

-lambda, lambda];

dk = opt.mod/gs.L(i)*k;

k = A(i)*opt.mod/gs.L(i)*k;

dk1v(:,i) = reshape(dk',[36,1]);

k_all(1:6,(6*i-5):(6*i)) = k;

Page 399: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

370

q = gs.LM(:,i);

K1(q,q) = K1(q,q) + k;

end

%% Unit Cell Periodicity constraints

K1_1 = K1;

% All node to node relations are based on this format

% NodeA - NodeB = NodeC - NodeD

% NA = NB + NC - ND

% For the stiffness matrix, in order to replace nodeA DOFs,

% col_B = col_B + col_A

% col_C = col_C + col_A

% col_D = col_D - col_A

corners = [ 25,105,121,125;

5, 5, 21,101;

21,101,101, 25;

1, 1, 1, 1];

Xedges = [ 6, 11, 16, 10, 15, 20,106,111,116;

1, 1, 1, 5, 5, 5,101,101,101;

110,115,120,110,115,120,110,115,120;

105,105,105,105,105,105,105,105,105];

Yedges = [ 2, 3, 4, 22, 23, 24,102,103,104;

1, 1, 1, 21, 21, 21,101,101,101;

122,123,124,122,123,124,122,123,124;

121,121,121,121,121,121,121,121,121];

Zedges = [ 26, 51, 76, 30, 55, 80, 46, 71, 96;

1, 1, 1, 5, 5, 5, 21, 21, 21;

50, 75,100, 50, 75,100, 50, 75,100;

25, 25, 25, 25, 25, 25, 25, 25, 25];

Xfaces = [ 27, 28, 29, 52, 53, 54, 77, 78, 79;

1, 1, 1, 1, 1, 1, 1, 1, 1;

47, 48, 49, 72, 73, 74, 97, 98, 99;

21, 21, 21, 21, 21, 21, 21, 21, 21];

Yfaces = [ 31, 36, 41, 56, 61, 66, 81, 86, 91;

1, 1, 1, 1, 1, 1, 1, 1, 1;

35, 40, 45, 60, 65, 70, 85, 90, 95;

5, 5, 5, 5, 5, 5, 5, 5, 5];

Zfaces = [ 7, 8, 9, 12, 13, 14, 17, 18, 19;

1, 1, 1, 1, 1, 1, 1, 1, 1;

107,108,109,112,113,114,117,118,119;

101,101,101,101,101,101,101,101,101];

nodes = [corners,Xedges,Yedges,Zedges,Xfaces,Yfaces,Zfaces];

for i = 1:length(nodes)

NA((3*i-2):3*i) = (3*nodes(1,i)-2):3*nodes(1,i);

NB((3*i-2):3*i) = (3*nodes(2,i)-2):3*nodes(2,i);

NC((3*i-2):3*i) = (3*nodes(3,i)-2):3*nodes(3,i);

ND((3*i-2):3*i) = (3*nodes(4,i)-2):3*nodes(4,i);

end

Page 400: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

371

for i = length(NA):-1:1

K1_1(:,NB(i)) = K1_1(:,NB(i)) + K1_1(:,NA(i));

K1_1(:,NC(i)) = K1_1(:,NC(i)) + K1_1(:,NA(i));

K1_1(:,ND(i)) = K1_1(:,ND(i)) - K1_1(:,NA(i));

end

% % Prescribed Displacements

pd1 = zeros(gs.Ndof,1);

pd2 = zeros(gs.Ndof,1);

pd3 = zeros(gs.Ndof,1);

pd4 = zeros(gs.Ndof,1);

pd5 = zeros(gs.Ndof,1);

pd6 = zeros(gs.Ndof,1);

nn = (1:gs.Ndof)';

INDEX = true(6,gs.Ndof); % All dof that will be removed

INDEX(:,NA) = false; % Set periodic dofs to be removed

PRDOF = zeros(6,gs.Ndof); % All dof with strain

% For prescribed dof's: displacement (e); Free (f); Fixed (0);

e = 1;

f = 2;

%% Case 1 - Axial X

% Prescribed DOFs

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,1) = [0;0;0]; % Grounded node

prenodes(:,5) = [0;f;0]; % Free in y

prenodes(:,21) = [e;0;0]; % Disp in x

prenodes(:,101) = [0;0;f]; % Free in z

prenodes(:,[ 47, 48, 49, 72, 73, 74, 97, 98, 99]) = repmat([e;f;f],[1,9]);

prenodes(:,[ 35, 40, 45, 60, 65, 70, 85, 90, 95]) = repmat([f;f;f],[1,9]);

prenodes(:,[107,108,109,112,113,114,117,118,119]) = repmat([f;f;f],[1,9]);

prenodes(:,[110,115,120]) = repmat([f;f;f],[1,3]);

prenodes(:,[ 50, 75,100]) = repmat([e;f;f],[1,3]);

prenodes(:,[122,123,124]) = repmat([e;f;f],[1,3]);

% Generate PRDOF showing strains and index for removing prescribed dofs

for i = 1:gs.N

for j = 1:3

dof = 3*(i-1)+j;

bc = prenodes(j,i);

if (bc == 0 || bc == e)

INDEX(1,dof) = false;

PRDOF(1,dof) = bc;

end

end

end

% Prepare Force Vector

F1 = -K1_1*PRDOF(1,:)';

% Reduce vectors and matrices based on prescribed and periodic

nn1 = nn(INDEX(1,:));

F1_r = F1(INDEX(1,:)); % Reduced Force vector

K1_1r = K1_1(INDEX(1,:),:);

K1_1r = K1_1r(:,INDEX(1,:));

%% Case 2 - Axial Y

Page 401: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

372

% Prescribed BC's

prenodes = f*ones(3,gs.N);

prenodes(:,5) = [0;e;0];

prenodes(:,1) = [0;0;0];

prenodes(:,21) = [f;0;0];

prenodes(:,101) = [0;0;f];

prenodes(:,[ 35, 40, 45, 60, 65, 70, 85, 90, 95]) = repmat([f;e;f],[1,9]);

prenodes(:,[ 47, 48, 49, 72, 73, 74, 97, 98, 99]) = repmat([f;f;f],[1,9]);

prenodes(:,[107,108,109,112,113,114,117,118,119]) = repmat([f;f;f],[1,9]);

prenodes(:,[ 50, 75,100]) = repmat([f;e;f],[1,3]);

prenodes(:,[110,115,120]) = repmat([f;e;f],[1,3]);

prenodes(:,[122,123,124]) = repmat([f;f;f],[1,3]);

% Generate PRDOF showing strains and index for removing prescribed dofs

for i = 1:size(prenodes,2)

for j = 1:3

dof = (3*(i-1) + j);

bc = prenodes(j,i);

if (bc == 0 || bc == e)

INDEX(2,dof) = false;

PRDOF(2,dof) = bc;

end

end

end

%Prepare Force Vector

F2 = -K1_1*(PRDOF(2,:)');

% Reduce vectors and matrices based on prescribed and periodic

nn2 = nn(INDEX(2,:));

F2_r = F2(INDEX(2,:)); % Reduced Force vector

K1_2r = K1_1(INDEX(2,:),:);

K1_2r = K1_2r(:,INDEX(2,:));

%% Case 3 - Axial Z

% Prescribed BC's

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,101) = [0;0;e]; % Strain in x, else fix

prenodes(:,1) = [0;0;0]; % All fixed

prenodes(:,5) = [0;f;0]; % All fixed

prenodes(:,21) = [f;0;0]; % All fixed

prenodes(:,[ 35, 40, 45, 60, 65, 70, 85, 90, 95]) = repmat([f;f;f],[1,9]); %

Fix y

prenodes(:,[ 47, 48, 49, 72, 73, 74, 97, 98, 99]) = repmat([f;f;f],[1,9]); %

Strain x

prenodes(:,[107,108,109,112,113,114,117,118,119]) = repmat([f;f;e],[1,9]); %

Fix z

prenodes(:,[ 50, 75,100]) = repmat([f;f;f],[1,3]); % Strain x, fix y

prenodes(:,[110,115,120]) = repmat([f;f;e],[1,3]); % Fix y,z

prenodes(:,[122,123,124]) = repmat([f;f;e],[1,3]); % Strain x, fix z

% Generate PRDOF showing strains and index for removing prescribed dofs

for i = 1:size(prenodes,2)

for j = 1:3

dof = (3*(i-1) + j);

bc = prenodes(j,i);

if (bc == 0 || bc == e)

INDEX(3,dof) = false;

PRDOF(3,dof) = bc;

Page 402: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

373

end

end

end

%Prepare Force Vector

F3 = -K1_1*(PRDOF(3,:)');

% Reduce vectors and matrices based on prescribed and periodic

nn3 = nn(INDEX(3,:));

F3_r = F3(INDEX(3,:)); % Reduced Force vector

K1_3r = K1_1(INDEX(3,:),:);

K1_3r = K1_3r(:,INDEX(3,:));

%% Case 4 - Shear YZ

% Prescribed BC's

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,5) = [0;0;e];

prenodes(:,1) = [0;0;0];

prenodes(:,101) = [0;0;f];

prenodes(:,21) = [f;0;0];

prenodes(:,[ 35, 40, 45, 60, 65, 70, 85, 90, 95]) = repmat([f;f;e],[1,9]);

prenodes(:,[ 47, 48, 49, 72, 73, 74, 97, 98, 99]) = repmat([f;f;f],[1,9]);

prenodes(:,[107,108,109,112,113,114,117,118,119]) = repmat([f;f;f],[1,9]);

prenodes(:,[ 50, 75,100]) = repmat([f;f;e],[1,3]);

prenodes(:,[110,115,120]) = repmat([f;f;e],[1,3]);

prenodes(:,[122,123,124]) = repmat([f;f;f],[1,3]);

% Generate PRDOF showing strains and index for removing prescribed dofs

for i = 1:size(prenodes,2)

for j = 1:3

dof = (3*(i-1) + j);

bc = prenodes(j,i);

if (bc == 0 || bc == e)

INDEX(4,dof) = false;

PRDOF(4,dof) = bc;

end

end

end

%Prepare Force Vector

F4 = -K1_1*(PRDOF(4,:)');

% Reduce vectors and matrices based on prescribed and periodic

nn4 = nn(INDEX(4,:));

F4_r = F4(INDEX(4,:)); % Reduced Force vector

K1_4r = K1_1(INDEX(4,:),:);

K1_4r = K1_4r(:,INDEX(4,:));

%% Case 5 - Shear

% Prescribed BC's

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,101) = [e;0;f]; % Strain in x, else fix

prenodes(:,1) = [0;0;0]; % All fixed

prenodes(:,5) = [0;f;0]; % All fixed

prenodes(:,21) = [f;0;0]; % All fixed

prenodes(:,[ 35, 40, 45, 60, 65, 70, 85, 90, 95]) = repmat([f;f;f],[1,9]); %

Fix y

prenodes(:,[ 47, 48, 49, 72, 73, 74, 97, 98, 99]) = repmat([f;f;f],[1,9]); %

Strain x

Page 403: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

374

prenodes(:,[107,108,109,112,113,114,117,118,119]) = repmat([e;f;f],[1,9]); %

Fix z

prenodes(:,[ 50, 75,100]) = repmat([f;f;f],[1,3]); % Strain x, fix y

prenodes(:,[110,115,120]) = repmat([e;f;f],[1,3]); % Fix y,z

prenodes(:,[122,123,124]) = repmat([e;f;f],[1,3]); % Strain x, fix z

% Generate PRDOF showing strains and index for removing prescribed dofs

for i = 1:size(prenodes,2)

for j = 1:3

dof = (3*(i-1) + j);

bc = prenodes(j,i);

if (bc == 0 || bc == e)

INDEX(5,dof) = false;

PRDOF(5,dof) = bc;

end

end

end

%Prepare Force Vector

F5 = -K1_1*(PRDOF(5,:)');

% Reduce vectors and matrices based on prescribed and periodic

nn5 = nn(INDEX(5,:));

F5_r = F5(INDEX(5,:)); % Reduced Force vector

K1_5r = K1_1(INDEX(5,:),:);

K1_5r = K1_5r(:,INDEX(5,:));

%% Case 6 - Shear

% Prescribed BC's

prenodes = f*ones(3,gs.N); % Start with all dof free

prenodes(:,21) = [f;e;0]; % Strain in x, else fix

prenodes(:,1) = [0;0;0]; % All fixed

prenodes(:,101) = [0;0;f]; % All fixed

prenodes(:,5) = [0;f;0]; % All fixed

prenodes(:,[ 35, 40, 45, 60, 65, 70, 85, 90, 95]) = repmat([f;f;f],[1,9]); %

Fix y

prenodes(:,[ 47, 48, 49, 72, 73, 74, 97, 98, 99]) = repmat([f;e;f],[1,9]); %

Strain x

prenodes(:,[107,108,109,112,113,114,117,118,119]) = repmat([f;f;f],[1,9]); %

Fix z

prenodes(:,[ 50, 75,100]) = repmat([f;e;f],[1,3]); % Strain x, fix y

prenodes(:,[110,115,120]) = repmat([f;f;f],[1,3]); % Fix y,z

prenodes(:,[122,123,124]) = repmat([f;e;f],[1,3]); % Strain x, fix z

% Generate PRDOF showing strains and index for removing prescribed dofs

for i = 1:size(prenodes,2)

for j = 1:3

dof = (3*(i-1) + j);

bc = prenodes(j,i);

if (bc == 0 || bc == e)

INDEX(6,dof) = false;

PRDOF(6,dof) = bc;

end

end

end

%Prepare Force Vector

F6 = -K1_1*(PRDOF(6,:)');

% Reduce vectors and matrices based on prescribed and periodic

Page 404: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

375

nn6 = nn(INDEX(6,:));

F6_r = F6(INDEX(6,:)); % Reduced Force vector

K1_6r = K1_1(INDEX(6,:),:);

K1_6r = K1_6r(:,INDEX(6,:));

%% Solve all test cases and rebuild displacement vectors

% Solve FEA

ua1_r = K1_1r\F1_r;

ua2_r = K1_2r\F2_r;

ua3_r = K1_3r\F3_r;

ua4_r = K1_4r\F4_r;

ua5_r = K1_5r\F5_r;

ua6_r = K1_6r\F6_r;

% Fill in prescribed/calculated dof

UA = [PRDOF(1,:);

PRDOF(2,:);

PRDOF(3,:);

PRDOF(4,:);

PRDOF(5,:);

PRDOF(6,:)];

% Fill in solved dof

UA(1,nn1) = ua1_r;

UA(2,nn2) = ua2_r;

UA(3,nn3) = ua3_r;

UA(4,nn4) = ua4_r;

UA(5,nn5) = ua5_r;

UA(6,nn6) = ua6_r;

% Calculate and fill in periodic displacements

for i = 1:length(NA)

UA(:,NA(i)) = UA(:,NB(i)) + UA(:,NC(i)) - UA(:,ND(i));

end

% Reshape to xy solutions

UA_XY = [reshape(UA(1,:),[3,gs.Ndof/3]);

reshape(UA(2,:),[3,gs.Ndof/3]);

reshape(UA(3,:),[3,gs.Ndof/3]);

reshape(UA(4,:),[3,gs.Ndof/3]);

reshape(UA(5,:),[3,gs.Ndof/3]);

reshape(UA(6,:),[3,gs.Ndof/3])];

% % Calculate new lengths

% newcoord = repmat(gs.coord,6,1) + UA_XY;

% for i = 1:gs.M

% nodeA = gs.IEN(1,i);

% nodeB = gs.IEN(2,i);

% pos1 = newcoord(1:3,nodeA);

% pos2 = newcoord(1:3,nodeB);

% gs.Ldef(i,:) = sqrt(sum((pos2-pos1).^2));

% end

%% Calculate Element Mutual Energies

D = zeros(36,gs.M);

% Prepare induced displacement matrices

for i = 1:gs.M

nodeA = gs.IEN(1,i);

nodeB = gs.IEN(2,i);

Page 405: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

376

D( 1:6 ,i) = [UA_XY(1:3,nodeA); UA_XY(1:3,nodeB)];

D( 7:12,i) = [UA_XY(4:6,nodeA); UA_XY(4:6,nodeB)];

D(13:18,i) = [UA_XY(7:9,nodeA); UA_XY(7:9,nodeB)];

D(19:24,i) = [UA_XY(10:12,nodeA); UA_XY(10:12,nodeB)];

D(25:30,i) = [UA_XY(13:15,nodeA); UA_XY(13:15,nodeB)];

D(31:36,i) = [UA_XY(16:18,nodeA); UA_XY(16:18,nodeB)];

end

%%%% Compute Elasticity Matrix

A_min = min(A);

A_rel = find(A>A_min);

Y = opt.ucs(1);

q = zeros(6);

for h = 1:gs.M

for i = 1:6

for j = 1:6

val = (1/Y)*(D((6*i-5):(6*i),h))'*...

k_all(:,(6*h-5):(6*h))*(D((6*j-5):(6*j),h));

q(i,j) = q(i,j) + val;

end

end

end

Eh = q;

%%% Calculate Buckling Load

Kbuck = 1;

F_crit = -

pi*pi*opt.mod*(Abuck(A_rel)'.*Abuck(A_rel)'/4/pi)./(Kbuck*gs.L(A_rel)).^2;

F_crit = gs.surfred(A_rel)'.*F_crit; % Account for surface

L_chng = (sum((D(16:18,A_rel)-D(13:15,A_rel)).^2).^(.5))'; % Deformed length

F_axi = Abuck(A_rel)'*opt.mod.*(L_chng)./gs.L(A_rel); % Axial force in each

member

factor2crit = -F_crit./F_axi;

F_eff = (opt.ucs(1)^2)*Eh(3,3)*(e)/opt.ucs(1);

min2buckle = F_eff*min(factor2crit(factor2crit>0));

% ALTO3D_plotdeformed(gs.Ainitial,gs,UA_XY(1:3,:)+gs.coord,A_relevant)

% ALTO3D_plotdeformed(gs.Ainitial,gs,UA_XY(4:6,:)+gs.coord,A_relevant)

% ALTO3D_plotdeformed(gs.Ainitial,gs,UA_XY(7:9,:)+gs.coord,A_relevant)

% ALTO3D_plotdeformed(gs.Ainitial,gs,UA_XY(10:12,:)+gs.coord,A_relevant)

% ALTO3D_plotdeformed(gs.Ainitial,gs,UA_XY(13:15,:)+gs.coord,A_relevant)

% ALTO3D_plotdeformed(gs.Ainitial,gs,UA_XY(16:18,:)+gs.coord,A_relevant)

end

Output Function

The output function is for the gradient-based optimization algorithm fmincon. It stores the

output information after each iteration such as the current unit cell design, the objective function

score, constraint violations, etc.

Page 406: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

377

ALTO3D_out.m

function stop = ALTO3D_out(xreduced,optimValues,state,opt)

stop = false;

iter = optimValues.iteration;

if ~isempty(opt.tempname)

load(opt.tempname)

end

switch state

case 'init'

hold on

case 'iter'

iter = iter + 1;

xreduced([optim.keep]) = xreduced;

xreduced([optim.remove]) = gs.Alower(1);

x_all = xreduced;

if ~isempty(optim.en_symmap)

for i = 1:gs.M

x_all(i) = xreduced(optim.en_symmap(i));

end

end

outvars(iter).FuncCount = optimValues.funccount;

outvars(iter).A = x_all';

outvars(iter).vol = x_all*gs.L;

outvars(iter).fval = optimValues.fval;

outvars(iter).Amid = nnz(x_all < ((pi/4*(opt.drel)^2)*ones(1,gs.M))

& x_all > 1e6*gs.Alower);

fprintf('%s \t %-7d %.8f \n',datestr(now,13),iter,optimValues.fval)

% Save output to file

if ~isempty(opt.tempname)

save(opt.tempname,'optim','outvars','-append')

end

case 'done'

iter = iter + 2;

xreduced([optim.keep]) = xreduced;

xreduced([optim.remove]) = gs.Alower(1);

x_all = xreduced;

if ~isempty(optim.en_symmap)

for i = 1:gs.M

x_all(i) = xreduced(optim.en_symmap(i));

end

end

outvars(iter).A = x_all';

outvars(iter).vol = x_all*gs.L;

outvars(iter).fval = optimValues.fval;

if opt.saveobj == 1

outvars(iter).objs = outvars(iter-1).objs(end,:);

outvars(iter).P = outvars(iter-1).P(end,:);

end

fprintf('%s \t %-7d %.8f \n',datestr(now,13),iter,optimValues.fval)

% Save output to file

if ~isempty(opt.tempname)

Page 407: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

378

save(opt.tempname,'optim','outvars','-append')

end

hold off

otherwise

end

end

GA Output Function

The genetic algorithm output function stores all the output information for the

evolutionary algorithm NSGA-II. At the end of each generation, it stores all the designs simulated

and their objective function scores. It also tracks the convergence or spread change.

ALTO3D_gaout.m

function [state,options,optchanged] = ALTO3D_gaout(options,state,flag,opt)

optchanged = false;

load(opt.filename)

switch flag

case {'init'}

disp(['Generation ',num2str(state.Generation),' Complete @

',datestr(now)])

outvars(state.Generation+1) = state;

save(opt.filename,'options','outvars','-append')

case {'iter'}

disp(['Generation ',num2str(state.Generation),' Complete @

',datestr(now)])

% outvars(state.Generation+1) = state;

outvars(2) = state;

save(opt.filename,'outvars','-append')

case {'done'}

end

end

nTopology Unit Cell Function

The nTopology unit cell function is for generating a file which can be directly imported

into nTop Platform as a lattice or thickened lattice. Using this function, the results of the

Page 408: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

379

optimized can be transferred into a representation of the 3D model that is created in nTopology

software.

ALTO3D_nToplat.m

function []=ALTO3D_nToplat(A,gs,tol,name)

filename = [name,'.ltcx'];

save(filename)

fileID = fopen(filename,'w');

fprintf(fileID,'<?xml version="1.0" encoding="UTF-8"?> \n<!--comment--> \n');

fprintf(fileID,['<graph id="0" name="',name,'" units="mm" type="rnd">\n']);

fprintf(fileID,'\t<nodegroup>\n');

offset = max(max(gs.coord)); % Used to place the centroid of the unit cell at

(0,0)

scale = 1;

nodeid = 0;

for i = 1:length(A)

if A(i)>tol

r = sqrt(A(i)/pi)/scale;

nodeA = gs.IEN(1,i);

nodeB = gs.IEN(2,i);

x1 = (gs.coord(1,nodeA) - 0.5*offset)/scale;

y1 = (gs.coord(2,nodeA) - 0.5*offset)/scale;

z1 = (gs.coord(3,nodeA) - 0.5*offset)/scale;

x2 = (gs.coord(1,nodeB) - 0.5*offset)/scale;

y2 = (gs.coord(2,nodeB) - 0.5*offset)/scale;

z2 = (gs.coord(3,nodeB) - 0.5*offset)/scale;

fprintf(fileID,'\t\t<node id="%d" x="%f" y="%f" z="%f"

r="%f"/>\n',nodeid,x1,y1,z1,r);

fprintf(fileID,'\t\t<node id="%d" x="%f" y="%f" z="%f"

r="%f"/>\n',(nodeid+1),x2,y2,z2,r);

% fprintf(fileID,'\t\t<node id="%d" x="%f" y="%f"

z="%f"/>\n',nodeid,x1,y1,z1);

% fprintf(fileID,'\t\t<node id="%d" x="%f" y="%f"

z="%f"/>\n',(nodeid+1),x2,y2,z2);

nodeid = nodeid + 2;

end

end

fprintf(fileID,'\t</nodegroup>\n');

fprintf(fileID,'\t<beamgroup>\n');

beamid = 0;

nodeid = 0;

for i = 1:length(A)

if A(i)>tol

fprintf(fileID,'\t\t<beam id="%d" n1="%d"

n2="%d"/>\n',beamid,nodeid,(nodeid+1));

beamid = beamid + 1;

nodeid = nodeid + 2;

end

end

fprintf(fileID,'\t</beamgroup>\n');

Page 409: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

380

fprintf(fileID,'</graph>\n');

fclose(fileID);

Plot Function

The plot function generates a 3D plot of the ground structure for analyzing results or

debugging.

ALTO3D_plot.m

% Plot Undeformed Truss

function ALTO3D_plot(A,IEN,coord,Arel)

% Plot Initial Truss

Amax = max(A);

Amax = 3.14;

Amin = min(A);

dim = size(coord,1);

if dim == 2

gap = 0;

x_lim = [min(coord(1,:))-gap, max(coord(1,:))+gap];

y_lim = [min(coord(2,:))-gap, max(coord(2,:))+gap];

figure

axis equal off tight

axis([x_lim y_lim])

xlabel('X Axis')

ylabel('Y Axis')

for i=1:length(A)

if (A(i) >= Arel)

w(i) = A(i)/(Amax)*10;

nodeA=IEN(1,i);

nodeB=IEN(2,i);

xA = coord(1,nodeA);

yA = coord(2,nodeA);

xB = coord(1,nodeB);

yB = coord(2,nodeB);

line([xA,xB],[yA,yB],'LineWidth',w(i))

end

end

elseif dim == 3

gap = 1;

x_lim = [min(coord(1,:))-gap, max(coord(1,:))+gap];

y_lim = [min(coord(2,:))-gap, max(coord(2,:))+gap];

z_lim = [min(coord(3,:))-gap, max(coord(3,:))+gap];

figure

axis manual equal

axis([x_lim y_lim z_lim])

Page 410: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

381

xlabel('X Axis')

ylabel('Y Axis')

zlabel('Z Axis')

for i=1:length(A)

if (A(i) >= Arel)

w(i) = A(i)/(44.2)*10;

nodeA=IEN(1,i);

nodeB=IEN(2,i);

xA = coord(1,nodeA);

yA = coord(2,nodeA);

zA = coord(3,nodeA);

xB = coord(1,nodeB);

yB = coord(2,nodeB);

zB = coord(3,nodeB);

line([xA,xB],[yA,yB],[zA,zB],'LineWidth',w(i))

end

end

else

error(['BH:Dimension is ',dim])

end

end

Plot Deformed Function

The plot deformed function is used to generate a 3D plot of the deformed ground

structure. The inputs are the initial and deformed positions of the nodes and then it recreates all

the elements.

ALTO3D_plotdeformed.m

% Plot Deformed Truss

function ALTO3D_plotdeformed(A,gs,coord_disp,A_relevant)

dim = size(coord_disp,1);

if dim == 2

% Plot Deformed Truss

figure

axis equal off tight

xlabel('X Axis')

ylabel('Y Axis')

for j=1:2

for i=1:gs.M

if A(i)>= A_relevant

if j==1

nodeA=gs.IEN(1,i);

nodeB=gs.IEN(2,i);

xA=gs.coord(1,nodeA);

yA=gs.coord(2,nodeA);

Page 411: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

382

xB=gs.coord(1,nodeB);

yB=gs.coord(2,nodeB);

line([xA,xB],[yA,yB],'Color','blue','LineWidth',3)

else

nodeA=gs.IEN(1,i);

nodeB=gs.IEN(2,i);

xA=coord_disp(1,nodeA);

yA=coord_disp(2,nodeA);

xB=coord_disp(1,nodeB);

yB=coord_disp(2,nodeB);

line([xA,xB],[yA,yB],'Color','red','LineWidth',3)

end

end

end

end

hold on

% plot(gs.coord(1,:),gs.coord(2,:),'*')

% plot(coord_disp(1,:),coord_disp(2,:),'*')

% % X Compression Arrows

% quiver(coord_disp(1,14)+.5,coord_disp(2,14),-

0.5,0,'Linewidth',3,'MaxHeadSize',2,'Color','red')

% quiver(coord_disp(1,15)+.5,coord_disp(2,15),-

0.5,0,'Linewidth',3,'MaxHeadSize',2,'Color','red')

% % Y Compression Arrows

% quiver(coord_disp(1, 8),coord_disp(2, 8)+0.5,0,-

0.5,'Linewidth',3,'MaxHeadSize',2,'Color','red')

% quiver(coord_disp(1,12),coord_disp(2,12)+0.5,0,-

0.5,'Linewidth',3,'MaxHeadSize',2,'Color','red')

elseif dim == 3

% Plot Deformed Truss

gap = 2;

x_lim = [min(gs.coord(1,:))-gap, max(gs.coord(1,:))+gap];

y_lim = [min(gs.coord(2,:))-gap, max(gs.coord(2,:))+gap];

z_lim = [min(gs.coord(3,:))-gap, max(gs.coord(3,:))+gap];

figure

axis manual equal

axis([x_lim y_lim z_lim])

xlabel('X Axis')

ylabel('Y Axis')

zlabel('Z Axis')

for j=1:2

for i=1:gs.M

if A(i)>= A_relevant

if j==1

nodeA=gs.IEN(1,i);

nodeB=gs.IEN(2,i);

xA=gs.coord(1,nodeA);

yA=gs.coord(2,nodeA);

zA=gs.coord(3,nodeA);

xB=gs.coord(1,nodeB);

yB=gs.coord(2,nodeB);

zB=gs.coord(3,nodeB);

line([xA,xB],[yA,yB],[zA,zB],'Color','blue','LineWidth',1)

else

nodeA=gs.IEN(1,i);

nodeB=gs.IEN(2,i);

Page 412: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

383

xA=coord_disp(1,nodeA);

yA=coord_disp(2,nodeA);

zA=coord_disp(3,nodeA);

xB=coord_disp(1,nodeB);

yB=coord_disp(2,nodeB);

zB=coord_disp(3,nodeB);

line([xA,xB],[yA,yB],[zA,zB],'Color','red','LineWidth',1)

end

end

end

end

grid ON

hold on

plot3(gs.coord(1,:),gs.coord(2,:),gs.coord(3,:),'*')

plot3(coord_disp(1,:),coord_disp(2,:),coord_disp(3,:),'r*')

% % X Compression Arrows

% quiver(coord_disp(1,14)+.5,coord_disp(2,14),-

0.5,0,'Linewidth',3,'MaxHeadSize',2,'Color','red')

% quiver(coord_disp(1,15)+.5,coord_disp(2,15),-

0.5,0,'Linewidth',3,'MaxHeadSize',2,'Color','red')

%

% % Y Compression Arrows

% quiver(coord_disp(1, 8),coord_disp(2, 8)+0.5,0,-

0.5,'Linewidth',3,'MaxHeadSize',2,'Color','red')

% quiver(coord_disp(1,12),coord_disp(2,12)+0.5,0,-

0.5,'Linewidth',3,'MaxHeadSize',2,'Color','red')

end

Standardized Unit Cell Simulation File

The standardized unit cell simulation file is for evaluating or creating models of generic

unit cells such as the BCC, FCC, etc. It can be used to generate and check specific unit cell types

for comparison against optimized results.

CellSim.m

close all;

clear;

% clc;

opt.type = 'interlocked'; % Lattice Type

opt.ucs = [8, 8, 8]; % Unit cell size in mm

opt.diam = [sqrt(2),1.75, 2, 2.25, 2.5, 3]; % Beam Diameter

opt.mod = 1.15e5; % Modulus of elasticity N/mm^2

opt.XYtol = []; % XY plane hatch offset for no contours

opt.layer = []; % Layer height accounts for stair step by creating area

to be ellipse

opt.new = 0; % 0 homog5, 1 homog5new, 2 homog5v2

opt.dlim = [9.5e-5, 9.5]; % Element diameter limits [min,max] in mm

opt.drel = [0.1]; % Relevant diameter (mm) - Used to determine which members

are relevant

Page 413: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

384

opt.adjarea = false; % Adjust cross section area to account for periodicity

opt.objs.type = 1;

opt.removenode = [];

prf = false; % Compute prf's?

opt.prfrel = 4;

A_relevant = (pi/4*(opt.drel)^2);

optim = struct('blank',[]);

switch lower(opt.type)

case{'sig8a'} % Correct

array = 4;

elems =

[21,205,703,851,1114,1117,1129,1159,1171,1276,1291,1329,1666,1669,1676,1695,170

4,1764,1782,1806];

case{'sig8d'} % Incorrect

array = 4;

elems =

[64,82,146,742,752,804,1074,1086,1200,1212,1252,1366,1642,1720,1761,1651,1727,1

825,1912,2014];

case{'cubic5all'}

array = 5;

elems =

[1,5,25,125,129,149,248,252,272,370,374,394,495,515,611,615,635,730,734,754,848

,852,872,965,969,989,1085,1105,1196,1200,1220,1310,1314,1334,1423,1427,1447,153

5,1539,1559,1650,1670,1756,1760,1780,1865,1869,1889,1973,1977,1997,2080,2084,21

04,2190,2210,2291,2315,2395,2419,2498,2522,2600,2624,2725,2801,2805,2825,2900,2

904,2924,2998,3002,3022,3095,3099,3119,3195,3215,3286,3290,3310,3380,3384,3404,

3473,3477,3497,3565,3569,3589,3660,3680,3746,3750,3770,3835,3839,3859,3923,3927

,3947,4010,4014,4034,4100,4120,4181,4185,4205,4265,4269,4289,4348,4352,4372,443

0,4434,4454,4515,4535,4591,4615,4670,4694,4748,4772,4825,4849,4925,4976,4980,50

00,5050,5054,5074,5123,5127,5147,5195,5199,5219,5270,5290,5336,5340,5360,5405,5

409,5429,5473,5477,5497,5540,5544,5564,5610,5630,5671,5675,5695,5735,5739,5759,

5798,5802,5822,5860,5864,5884,5925,5945,5981,5985,6005,6040,6044,6064,6098,6102

,6122,6155,6159,6179,6215,6235,6266,6290,6320,6344,6373,6397,6425,6449,6500,652

6,6530,6550,6575,6579,6599,6623,6627,6647,6670,6674,6694,6720,6740,6761,6765,67

85,6805,6809,6829,6848,6852,6872,6890,6894,6914,6935,6955,6971,6975,6995,7010,7

014,7034,7048,7052,7072,7085,7089,7109,7125,7145,7156,7160,7180,7190,7194,7214,

7223,7227,7247,7255,7259,7279,7290,7310,7316,7340,7345,7369,7373,7397,7400,7424

,7450,7451,7455,7475,7479,7498,7502,7520,7524,7545,7561,7565,7580,7584,7598,760

2,7615,7619,7635,7646,7650,7660,7664,7673,7677,7685,7689,7700,7706,7710,7715,77

19,7723,7727,7730,7734,7740,7741,7745,7748,7750];

case{'homogerr'}

array = 5;

elems = [4987,5799,5805]; % Middle

% elems = [ 12,1424,1430]; %10,1205 % 1197

case{'homogerr2x2'}

array = 5;

% elems =

[2806,3003,3380,3383,3565,3568,3751,3928,4265,4268,4430,4433,6531,6628,6805,680

8,6890,6893,6976,7053,7190,7193,7255,7258]; % Middle

elems =

[6,253,730,733,965,968,1201,1428,1865,1868,2080,2083,4981,5128,5405,5408,5540,5

543,5676,5803,6040,6043,6155,6158];

case{'homogerr20'}

array = 5;

elems =

[6,253,730,733,965,968,1201,1428,1865,1868,2080,2083,2806,3003,3380,3383,3565,3

568,3751,3928,4265,4268,4430,4433,4981,5128,5405,5408,5540,5543,5676,5803,6040,

6043,6155,6158,6531,6628,6805,6808,6890,6893,6976,7053,7190,7193,7255,7258,7456

,7503,7580,7583,7615,7618,7651,7678,7715,7718,7730,7733];

Page 414: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

385

case{'cubic'}

array = 4;

elems = [ 3,12,48,198,234,693,738,888,1899,1908,1950,2013];

case{'plus'}

array = 5;

elems = [ 1472,5132,5672,5799,5807,5847];

case{'plus2x2'}

array = 5;

elems =

[754,989,1889,2104,2904,3099,3286,3380,3384,3404,3473,3565,3569,3589,3839,4014,

4181,4265,4269,4289,4348,4430,4434,4454,5429,5564,6064,6179,6579,6674,6761,6805

,6809,6829,6848,6890,6894,6914,7014,7089,7156,7190,7194,7214,7223,7255,7259,727

9];

case{'cubic5'}

array = 5;

elems = [ 4,20,100,510,590,2294,2390,2800,7454,7470,7560,7744];

case{'cubic52x2'}

array = 5;

elems =

[2,10,50,249,257,297,500,540,1197,1205,1245,1424,1432,1472,1655,1695,2292,2340,

2499,2547,2750,4977,4985,5025,5124,5132,5172,5275,5315,5672,5680,5720,5799,5807

,5847,5930,5970,6267,6315,6374,6422,6525,7452,7460,7499,7507,7550,7647,7655,767

4,7682,7705,7742,7749];

case{'bcc'}

array = 5;

elems = [62,548,2332,2738,5835,5839,5855,5859];

case{'bcc2x2'}

array = 5;

elems =

[31,276,278,519,1216,1226,1441,1443,1451,1453,1664,1674,2311,2516,2518,2719,339

8,3400,3408,3410,3583,3585,3593,3595,4283,4285,4293,4295,4448,4450,4458,4460,50

06,5151,5153,5294,5691,5701,5816,5818,5826,5828,5939,5949,6286,6391,6393,6494,6

823,6825,6833,6835,6908,6910,6918,6920,7208,7210,7218,7220,7273,7275,7283,7285]

;

case{'bccz'}

array = 5;

elems = [62,100,548,590,2332,2390,2738,2800,5835,5839,5855,5859];

case{'bccz2x2'}

array = 5;

elems =

[31,50,276,278,297,519,540,1216,1226,1245,1441,1443,1451,1453,1472,1664,1674,16

95,2311,2340,2516,2518,2547,2719,2750,3398,3400,3408,3410,3583,3585,3593,3595,4

283,4285,4293,4295,4448,4450,4458,4460,5006,5025,5151,5153,5172,5294,5315,5691,

5701,5720,5816,5818,5826,5828,5847,5939,5949,5970,6286,6315,6391,6393,6422,6494

,6525,6823,6825,6833,6835,6908,6910,6918,6920,7208,7210,7218,7220,7273,7275,728

3,7285];

case{'fcc'}

array = 5;

elems =

[12,52,60,498,538,550,1430,1434,2330,2342,2740,2748,5170,5174,5710,5730,5960,59

80,6420,6424,7462,7548,7680,7684];

case{'fccnohrz'}

array = 5;

elems =

[52,60,538,550,2330,2342,2740,2748,5170,5174,5710,5730,5960,5980,6420,6424];

case{'fccz'}

array = 5;

elems =

[12,52,60,100,498,538,550,590,1430,1434,2330,2342,2390,2740,2748,2800,5170,5174

,5710,5730,5960,5980,6420,6424,7462,7548,7680,7684];

case{'fccznohrz'}

Page 415: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

386

array = 5;

elems =

[52,60,100,538,550,590,2330,2342,2390,2740,2748,2800,5170,5174,5710,5730,5960,5

980,6420,6424];

case{'f2bcc'}

array = 5;

elems =

[12,52,60,498,538,550,1430,1434,2330,2342,2740,2748,5170,5174,5710,5730,5960,59

80,6420,6424,7462,7548,7680,7684,62,548,2332,2738,5835,5839,5855,5859];

case{'f2bccz'}

array = 5;

elems =

[62,100,548,590,2332,2390,2738,2800,5835,5839,5855,5859,12,52,60,498,538,550,14

30,1434,2330,2342,2740,2748,5170,5174,5710,5730,5960,5980,6420,6424,7462,7548,7

680,7684];

case{'f2bccnohrz'}

array = 5;

elems =

[62,548,2332,2738,5835,5839,5855,5859,52,60,538,550,2330,2342,2740,2748,5170,51

74,5710,5730,5960,5980,6420,6424];

case{'f2bccznohrz'}

array = 5;

elems =

[62,548,2332,2738,5835,5839,5855,5859,52,60,100,538,550,590,2330,2342,2390,2740

,2748,2800,5170,5174,5710,5730,5960,5980,6420,6424];

case{'column'}

array = 5;

elems = [1522];

case{'columns'}

array = 5;

elems = [100,590,1522,2390,2800];

case{'diamond'}

array = 5;

% Connected to 1

% elems =

[31,1441,1453,2719,3400,3408,4450,4458,5153,5701,5939,6391,6910,6918,7210,7218]

;

% Connected to 5

elems =

[519,1443,1451,2311,3583,3595,4283,4295,5151,5691,5949,6393,6823,6835,7273,7285

];

case{'interlocked'} % Same as 8-1 Diamond with sqrt(2) as diameter

array = 5;

% Connected to 1

% elems = [62,2738,5839,5855];

% Connected to 5

elems = [548,2332,5835,5859];

case{'interlocked2x2'} % Interlocked 2x2 homogenization

array = 5;

% All

elems =

[31,278,1226,1441,1453,1664,2516,2719,3400,3408,3585,3593,4285,4293,4450,4458,5

006,5153,5701,5816,5828,5939,6391,6494,6825,6833,6910,6918,7210,7218,7275,7283]

;

% One set

% elems =

[31,1441,1453,2719,3400,3408,4450,4458,5153,5701,5939,6391,6910,6918,7210,7218]

;

% Second set

Page 416: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

387

% elems =

[278,1226,1664,2516,3585,3593,4285,4293,5006,5816,5828,6494,6825,6833,7275,7283

];

case{'fluorite'}

array = 5;

elems =

[31,519,1441,1443,1451,1453,2311,2719,3400,3408,3583,3595,4283,4295,4450,4458,5

151,5153,5691,5701,5939,5949,6391,6393,6823,6835,6910,6918,7210,7218,7273,7285]

;

case{'octet'}

array = 5;

elems =

[12,52,60,498,538,550,1430,1434,1462,1470,1474,1482,2330,2342,2740,2748,5130,51

34,5170,5174,5182,5682,5710,5722,5730,5928,5960,5968,5980,6412,6420,6424,7462,7

548,7680,7684];

case{'octet2x2'}

array = 5;

elems =

[6,26,30,251,253,271,273,277,494,514,520,733,735,749,753,755,759,968,970,984,98

8,990,994,1201,1215,1221,1225,1426,1428,1442,1446,1448,1452,1649,1665,1669,1675

,1868,1870,1884,1888,1890,1894,2083,2085,2099,2103,2105,2109,2310,2316,2517,252

1,2523,2720,2724,2903,2905,2923,2925,2929,3098,3100,3118,3120,3124,3291,3305,33

11,3315,3476,3478,3492,3496,3498,3502,3659,3675,3679,3685,3838,3840,3854,3858,3

860,3864,4013,4015,4029,4033,4035,4039,4186,4200,4206,4210,4351,4353,4367,4371,

4373,4377,4514,4530,4534,4540,4689,4693,4695,4844,4848,4850,4981,5001,5005,5126

,5128,5146,5148,5152,5269,5289,5295,5408,5410,5424,5428,5430,5434,5543,5545,555

9,5563,5565,5569,5676,5690,5696,5700,5801,5803,5817,5821,5823,5827,5924,5940,59

44,5950,6043,6045,6059,6063,6065,6069,6158,6160,6174,6178,6180,6184,6285,6291,6

392,6396,6398,6495,6499,6578,6580,6598,6600,6604,6673,6675,6693,6695,6699,6766,

6780,6786,6790,6851,6853,6867,6871,6873,6877,6934,6950,6954,6960,7013,7015,7029

,7033,7035,7039,7088,7090,7104,7108,7110,7114,7161,7175,7181,7185,7226,7228,724

2,7246,7248,7252,7289,7305,7309,7315,7364,7368,7370,7419,7423,7425,7456,7501,75

03,7544,7583,7585,7618,7620,7651,7676,7678,7699,7718,7720,7733,7735];

case{'truncated cube'}

array = 4;

elems =

[64,66,78,130,142,250,258,424,432,481,496,641,658,742,756,808,904,1042,1402,150

4,1537,1540,1625,1630,1838,1843,1892,1895,1912,1914,1930,1954,1984,1993,2009,20

14];

case{'kelvin'}

array = 5;

elems =

[851,853,867,1315,1333,1538,1560,2002,3021,3023,3765,3775,4115,4125,4771,4773,5

053,5075,5200,5218,5365,5635,5986,6000,6214,6230,6345,6448,6652,6996,7144,7392,

7601,7603,7665,7688];

case{'isotruss'}

array = 5;

elems =

[4,20,62,100,510,548,590,1472,2294,2332,2390,2738,2800,5132,5672,5799,5807,5835

,5839,5847,5855,5859,7454,7470,7560,7744];

case{'4cols'}

array = 5;

elems = [829,1064,1964,2179];

case{'octahedral'} % Same as Octet(8,1) with diam = 1.41421

array = 5;

elems = [1462,1470,1474,1482,5130,5134,5182,5682,5722,5928,5968,6412];

case{'octahedral2'}

array = 5;

elems =

[255,259,295,299,1207,1235,1255,1653,1685,1705,2545,2549,5027,5035,5313,5325,63

05,6317,6515,6523,7505,7509,7657,7703];

Page 417: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

388

case{'octet2_15'} % Optioctet % Same as Octet(8,1) with perpendiculars =

1.189, diags = 1.3*Area_perps

array = 5;

elems =

[62,548,1472,2332,2738,5132,5672,5799,5807,5835,5839,5847,5855,5859];

case{'opti'} % Optioctet v2, Same as Octet(8,1) with perpendiculars =

1.189, diags = 1.3*Area_perps

array = 5;

elems =

[ 4,20,100,510,590,2294,2390,2800,7454,7470,7560,7744,62,548,2332,2738,5835,583

9,5855,5859];

case{'opti2x2'} % Optioctet 2x2 homogenization

array = 5;

elems =

[2,10,50,249,257,297,500,540,1197,1205,1245,1424,1432,1472,1655,1695,2292,2340,

2499,2547,2750,4977,4985,5025,5124,5132,5172,5275,5315,5672,5680,5720,5799,5807

,5847,5930,5970,6267,6315,6374,6422,6525,7452,7460,7499,7507,7550,7647,7655,767

4,7682,7705,7742,7749];

elems2=

[31,276,278,519,1216,1226,1441,1443,1451,1453,1664,1674,2311,2516,2518,2719,339

8,3400,3408,3410,3583,3585,3593,3595,4283,4285,4293,4295,4448,4450,4458,4460,50

06,5151,5153,5294,5691,5701,5816,5818,5826,5828,5939,5949,6286,6391,6393,6494,6

823,6825,6833,6835,6908,6910,6918,6920,7208,7210,7218,7220,7273,7275,7283,7285]

;

case{'column2x2'}

array = 5;

elems = [779,1014,1914,2129,5454,5589,6089,6204];

otherwise

error('Unknown element type')

end

%%

opt.gssize = [ array, array, array];

index = 1;

for i = [opt.diam]

for j = [opt.ucs(1)]

opt.diam = i;

opt.ucs = opt.ucs;

opt.init = pi/4*(opt.dlim(1)^2)*ones(1,((array^3)*(array^3-1)/2));

opt.init(elems) = pi/4*(opt.diam^2);

% Change some elems

if opt.type == "octet2_15"

% opt.init(elems([1,2,4,5,10,11,13,14])) = 1.3*pi/4*(opt.diam^2); %

For octet2_15

opt.init(elems([1,2,4,5,10,11,13,14])) =

pi/4*(sqrt(1.3)*opt.diam)^2; % For octet2_15

elseif opt.type == "opti"

opt.init([62,548,2332,2738,5835,5839,5855,5859]) =

1.3*pi/4*(opt.diam^2); % For cubic5bcc

% opt.init([62,548,2332,2738,5835,5839,5855,5859]) =

pi/4*(sqrt(1.3)*1.189/2-.05)^2; % For manufacturing error

elseif opt.type == "opti2x2"

opt.init(elems2) = 1.3*pi/4*(opt.diam^2);

end

gs = ALTO3D_initgs(opt);

A = gs.Ainitial;

if ~isempty(opt.XYtol) || ~isempty(opt.layer)

rel = (A>=A_relevant);

Page 418: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

389

R1 = sqrt(A(rel)/pi); % Always the radius of the cross section

parallel to build plate

R2 = R1; % R2 angle relative to build plate changes with build

angle

% A(rel) = pi/4*((sqrt((4/pi)*A(rel))-

(2*opt.XYtol)*sin(gs.thetab(rel)')).^2);

if ~isempty(opt.XYtol) % Contour Offset

R1 = R1 - opt.XYtol;

R2 = R2 - opt.XYtol*sin(gs.thetab(rel)');

end

if ~isempty(opt.layer) % Staircase Effect

BArel = gs.thetab(rel);

SCBArel = (BArel' > 0);

R2(SCBArel) = R2(SCBArel) - (opt.layer/2)*cos(BArel(SCBArel)');

end

% if ~isempty(opt.layer)

% R2 = R2 - opt.layer/2*cos(gs.thetab(rel)');

% end

A(rel) = pi*R1.*R2;

gs.Ainitial = A;

end

% if ~isempty(opt.layer)

% rel = (A>=A_relevant & gs.thetab'>0);

% A(rel) = pi*(sqrt(A(rel)/pi)).*(sqrt(A(rel)/pi)-

opt.layer/2*cos(gs.thetab(rel)')); %Ellipsoidal

% % A(rel) = pi/4*((sqrt((4/pi)*A(rel))-

(opt.layer*cos(gs.thetab(rel)'))).^2); % Circular

% gs.Ainitial = A;

% end

if ~opt.adjarea

A_notred = A;

A_surfred = A;

A_surfred(A>=A_relevant) =

gs.surfred(A>=A_relevant).*(A(A>=A_relevant));

Volfrac = (A_surfred*gs.L)/(opt.ucs(1)*opt.ucs(2)*opt.ucs(3))

if array == 4

[Eh,min2buckle] = ALTO3D_homog(A_surfred,A,opt,optim,gs);

elseif array == 5

if opt.new == 2

[Eh,min2buckle] =

ALTO3D_homog5v2(A_surfred,A,opt,optim,gs);

elseif opt.new == 1

[Eh,min2buckle] =

ALTO3D_homog5new(A_surfred,A,opt,optim,gs);

elseif opt.new == 0

[Eh,min2buckle] = ALTO3D_homog5(A_surfred,A,opt,optim,gs);

[Eherr,min2buckleerr] =

ALTO3D_homog5err(A_surfred,A,opt,optim,gs);

Eherr

val = 3.9917e3;

target = [val , val/2, val/2, 0, 0, 0;

val/2, val , val/2, 0, 0, 0;

val/2, val/2, val, 0, 0, 0;

0, 0, 0,val/2, 0, 0;

0, 0, 0, 0,val/2, 0;

0, 0, 0, 0, 0,val/2];

Page 419: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

390

aerr = abs(((target - Eherr)./target));

a = abs(((target - Eh)./target));

objerr = sum(aerr(aerr<inf));

obj = sum(a(a<inf));

else

error('Specify opt.new')

end

end

Eh

E33(index,1) = Eh(3,3)

M2B(index,1) = min2buckle;

GXY(index,1) = Eh(6,6)/2;

end

% PRF scores

if prf

if ~exist('middist','var')

% Perform in GS to find all midpoints and distances

% Find all midpoints

midpts = zeros(3,gs.M);

for k = 1:gs.M

midpts(:,k) =

0.5*gs.coord(:,gs.IEN(1,k))+0.5*gs.coord(:,gs.IEN(2,k));

end

% Calculate distance between all midpoints

middist = 1000*ones(gs.M,gs.M);

for k = 1:gs.M

for l = (k+1):gs.M

middist(k,l) = norm(midpts(:,l)-midpts(:,k));

end

end

end

% Realtime computation required to reduce and adjust distances for

% areas, make sure this does not apply to surface elements that

% have been reduced or it will be inaccurate.

A_rel = A(A>(pi/4*opt.drel^2));

d_rel = (4*A_rel/pi).^.5;

middist_red = middist(A>=(pi/4*opt.drel^2),A>=(pi/4*opt.drel^2));

middist_adj = middist_red;

for k = 1:length(A_rel)

for l = (k+1):length(A_rel)

middist_adj(k,l) = middist_red(k,l)-

0.5*(d_rel(k)+d_rel(l));

end

end

prfmin(index,1) = min(min(middist_adj(middist_adj<1000)))

prfavg(index,1) = mean(mean(middist_adj(middist_adj<1000)))

prfcnt = nnz(middist_adj < opt.prfrel)

end

% Connectivity

P=[];

A_rel = A>=A_relevant;

rel_nodes = unique(gs.IEN(:,(A_rel)'));

Page 420: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

391

check1 = sum(ismember(gs.edg_cor,rel_nodes),2);

check2 = sum(ismember(gs.faces,rel_nodes),2);

if any([check1;check2]==1)

flaws = nnz([check1;check2]==1)

P = [P,100];

end

index = index + 1;

end

end

ALTO3D_plot(gs.Ainitial,gs.IEN,gs.coord,A_relevant)

Page 421: DESIGN OPTIMIZATION OF COMPLIANT STRUCTURES FOR ...

VITA

BRADLEY HANKS

EDUCATION The Pennsylvania State University Doctor of Philosophy, Mechanical Engineering

Dissertation: “Design Optimization of Compliant Structures for Radiofrequency Ablation and Additive Manufacturing”

Minor in Additive Manufacturing and Design Master of Science, Mechanical Engineering

2020

2020 2018

Brigham Young University Bachelor of Science, Mechanical Engineering

2015

JOURNAL PUBLICATIONS Hanks, B., Berthel, J., Frecker, M., Simpson, T., 2020, “Mechanical Properties of Additively Manufactured Metal Lattice Structures: Data Review and Design Interface,” Additive Manufacturing, 35, p. 101301

2020

Hanks, B., Frecker, M., and Moyer, M., 2018, “Optimization of an Endoscopic Radiofrequency Ablation Electrode,” ASME J. Med. Devices, 12(3), pp. 1–11

2018

Sessions, J. W., Lindstrom, D. L., Hanks, B. W., Hope, S., and Jensen, B. D., 2016, “The Effect of Lance Geometry and Carbon Coating of Silicon Lances on Propidium Iodide Uptake in Lance Array Nanoinjection of HeLa 229 Cells,” J. Micromechanics Microengineering, 26(4).

2016

Sessions, J. W., Hanks, B. W., Lindstrom, D. L., Hope, S., and Jensen, B. D., 2016, “Transient Low Temperature Effects on Propidium Iodide Uptake in Lance Array Nanoinjected HeLa Cells,” J. Nanotechnol. Eng. Med., 6(4).

2016

Sessions, J. W., Skousen, C. S., Price, K. D., Hanks, B. W., Hope, S., Alder, J. K., and Jensen, B. D., 2016, “CRISPR-Cas9 Directed Knock-out of a Constitutively Expressed Gene Using Lance Array Nanoinjection,” Springerplus, 5(1), p. 1521.

2016

SELECTED CONFERENCE PUBLICATIONS Hanks, B., Frecker, M., 2020, “3D Additive Lattice Topology Optimization: A Unit Cell Design Approach”, ASME IDETC & CIE, DETC2020-19687, Accepted for publication

2020

Hanks, B., and Frecker, M., 2019, “Lattice Structure Design for Additive Manufacturing: Unit Cell Topology Optimization,” ASME IDETC & CIE, DETC2019-97863.

2019

Hanks, B., Azhar, F., Frecker, M., Clement, R., Greaser, J., and Snook, K., 2019, “A Deployable Multi-Tine Endoscopic Radiofrequency Ablation Electrode: Simulation Validation in a Thermochromic Tissue Phantom,” Design of Medical Devices Conference, DMD2019-3214.

2019

Hanks, B., Frecker, M., and Moyer, M., 2017, “Optimization of an Endoscopic Radiofrequency Ablation Electrode,” ASME IDETC & CIE, DETC2017- 673576

2017

HONORS AND AWARDS Marley Graduate Fellowship in Engineering – The Pennsylvania State University 2019 Reiss Fellowship in Engineering – The Pennsylvania State University 2019 1st Place Millennium Science Café Pitch Competition 2017 University Graduate Fellowship – The Pennsylvania State University 2015


Recommended