Supplementary Materials

Similar documents
Accepted Manuscript. 3D nano-architected metallic glass: Size effect suppresses catastrophic failure. Rachel Liontas, Julia R.

HONEYCOMB MECHANICAL BEHAVIOR USING MACROINDENTATION

MECHANICAL PROPERTIES AND TESTS. Materials Science

YIELD & TENSILE STRENGTH OF STEEL & ALUMINIUM USING MICROINDENTATION

RE-EXAMINATION OF NIST ACOUSTIC EMISSION SENSOR CALIBRATION: Part I Modeling the loading from glass capillary fracture

CHAPTER 7 ANALYTICAL PROGRAMME USING ABAQUS

Chapter 4 MECHANICAL PROPERTIES OF MATERIAL. By: Ardiyansyah Syahrom

To have a clear idea about what really happened and to prevent the

EDGE CHIPPING RESISTANCE USING MACROINDENTATION TESTING

Chapter 7. Finite Elements Model and Results

MSE 3143 Ceramic Materials

ELASTIC-PLASTIC RESPONSE OF POLYMER COATING UNDER REPETITIVE IMPACT

M3 Review Automated Nanoindentation

DCB TEST SAMPLE DESIGN FOR MICRO-MECHANICAL TESTING

Chapter 7: Mechanical Properties 1- Load 2- Deformation 3- Stress 4- Strain 5- Elastic behavior

INELASTIC SEISMIC RESPONSE ANALYSES OF REINFORCED CONCRETE BRIDGE PIERS WITH THREE-DIMENSIONAL FE ANALYSIS METHOD. Guangfeng Zhang 1, Shigeki Unjoh 2

Welcome to ENR116 Engineering Materials. This lecture summary is part of module 2, Material Properties.

Micro-Electro-Mechanical Systems (MEMS) Fabrication. Special Process Modules for MEMS. Principle of Sensing and Actuation

Evaluation of Mechanical Properties of Hard Coatings

In-situ Electron Microscopy Mechanical Testing for Steels

NANOINDENTATION CREEP MEASUREMENT

Supplementary Figures

Failure criterion of silver nanowire electrodes on a polymer substrate for highly flexible devices

CE 221: MECHANICS OF SOLIDS I CHAPTER 3: MECHANICAL PROPERTIES OF MATERIALS

Buckling Analysis of Thin Walled Cylinders Subjected to Axially Compressive Load by ANSYS

ELASTIC AND ELASTO-PLASTIC BUCKLING ANALYSIS OF PERFORATED STEEL PLATES

SEISMIC BEHAVIOR OF STEEL RIGID FRAME WITH IMPERFECT BRACE MEMBERS

NANOINDENTATION OF SILICON CARBIDE WAFER COATINGS

7 LOCAL BUCKLING OF STEEL CLASS 4 SECTION BEAMS

Measurement of Residual Stress by X-ray Diffraction

FME201 Solid & Structural Mechanics I Dr.Hussein Jama Office 414

CONTROLLED HUMIDITY NANOINDENTATION OF POLYMER FILMS

Finite Element Analysis on the Unloading Elastic Modulus of Aluminum Foams by Unit-cell Model

Effect of dual-size cell mix on the stiffness and strength of open-cell aluminum foams

Fracture. Brittle vs. Ductile Fracture Ductile materials more plastic deformation and energy absorption (toughness) before fracture.

Journal of Asian Scientific Research EVALUATION OF RECTANGULAR CONCRETE-FILLED STEEL-HOLLOW SECTION BEAM-COLUMNS

Multiphoton lithography based 3D micro/nano printing Dr Qin Hu

Effects of Film Thickness on the Yielding Behavior of Polycrystalline Gold Films

NUMERICAL SIMULATION OF CIRCULAR HOLLOW STEEL COLUMNS CONFINED WITH FIBER REINFORCED POLYMER UNDER AXIAL COMPRESSION

CHAPTER 5 FINITE ELEMENT MODELLING

BFF1113 Engineering Materials DR. NOOR MAZNI ISMAIL FACULTY OF MANUFACTURING ENGINEERING

SET PROJECT STRUCTURAL ANALYSIS OF A TROUGH MODULE STRUCTURE, IN OPERATION AND EMERGENCY Luca Massidda

When an axial load is applied to a bar, normal stresses are produced on a cross section perpendicular to the axis of the bar.

Supporting Information

Nonlinear Finite Element Modeling & Simulation

PSEUDO-DYNAMIC TEST AND NON-LINEAR ANALYSIS OF A 1:10 SCALE PRE-STRESSED CONCRETE CONTAIN VESSEL MODEL FOR CNP1000 NUCLEAR POWER PLANT

3. MECHANICAL PROPERTIES OF STRUCTURAL MATERIALS

Nano Puncture Resistance Using Nanoindentation

LRFD AND EUROCODE-3 SPECIFICATIONS FOR ULTIMATE LOAD CARRYING CAPACITY EVALUATION OF STEEL COLUMNS AND EFFECTS OF IMPERFECTIONS

FINITE ELEMENT MODELING OF ANGLE BRACING MEMBER BEHAVIOR IN EXPERIMENTALLY TESTED SUB-FRAME SPECIMENS

5. A round rod is subjected to an axial force of 10 kn. The diameter of the rod is 1 inch. The engineering stress is (a) MPa (b) 3.

Leelachai M, Benson S, Dow RS. Progressive Collapse of Intact and Damaged Stiffened Panels.

Theoretical Analysis of the Mechanical Response of Lattice Structures Manufactured using Selective Laser Melting

Modeling of friction and structural transformations in diamond-like carbon coatings

CHAPTER 3 OUTLINE PROPERTIES OF MATERIALS PART 1

INTERNATIONAL JOURNAL OF CIVIL AND STRUCTURAL ENGINEERING Volume 2, No 2, 2011

DIRECT MEASUREMENT ON FRACTURE TOUGHNESS OF CARBON FIBER ABSTRACT

factured pillars, even though the strength is significantly higher than in the bulk. These yield stress values, y

MAE 456 FINITE ELEMENT ANALYSIS Truss Buckling Analysis LAB INSTRUCTIONS LAB ASSIGNMENT 3. Lab Objectives. Lab Tasks

Reduced Ductility due to Local Variation in Material Properties for 3D-printed Components

Engineering Materials

ADVANCES in NATURAL and APPLIED SCIENCES

Chapter 7. Mechanical properties 7.1. Introduction 7.2. Stress-strain concepts and behaviour 7.3. Mechanical behaviour of metals 7.4.

Micro-Electro-Mechanical Systems (MEMS) Fabrication. Special Process Modules for MEMS. Principle of Sensing and Actuation

THE ENERGY ABSORPTION OF ALUMINIUM HONEYCOMB UNDER QUASI-STATIC LOADING

CHARACTERIZATION OF STOCHASTIC HONEYCOMB SANDWICH FAILURE

Punching shear resistance of reinforced concrete slabs under pneumatic loads

Supplementary Materials for

We are IntechOpen, the world s leading publisher of Open Access books Built by scientists, for scientists. International authors and editors

Increased resistance to buckling of piston rods through induction hardening

Metals are generally ductile because the structure consists of close-packed layers of

Compressive strength of double-bottom under alternate hold loading condition

CORROSION HARDNESS EFFECT OF STEEL SCREWS BY NANOINDENTATION

Mechanical Behavior and Characterization of Stern-shaft Mechanical Sealing Device. Yongjin Lu a *, Rui Lin

When an axial load is applied to a bar, normal stresses are produced on a cross section perpendicular to the axis of the bar.

MEMS 487. Class 04, Feb. 13, K.J. Hemker

New approach to improving distortional strength of intermediate length thin-walled open section columns

Multiscale Modeling of Metallic Materials Containing Embedded Particles

The Use of Sustainable Materials for Quick Repair of Aging Bridges

CHAPTER 5 FINITE ELEMENT MODELING

Citation for published version (APA): Amsterdam, E. (2008). Structural performance and failure analysis of aluminium foams s.n.

A practical appraisal method for through-wall cracks in tubular structures

J. Y. He, Z. L. Zhang, H. Kristiansen, Cracking and delamination of nanoscale metal coating on composite polymer particles. Proceedings of the 18 th

Numerical Study on Response of Laterally Loaded Piles in Soils

Study on ultimate strength of ship plates with calculated weld-induced residual stress

Ultimate Strength Analysis of Stiffened Panels Subjected to Biaxial Thrust Using JTP and JBP Methods

A Coupled Thermal and Mechanical Model of Sliding Wear

In situ TEM Characterization of Shear Stress-Induced Interlayer. Sliding in the Cross Section View of Molybdenum Disulfide

Compressive strength of double-bottom under alternate hold loading condition

a. 50% fine pearlite, 12.5% bainite, 37.5% martensite. 590 C for 5 seconds, 350 C for 50 seconds, cool to room temperature.

THE INFLUENCE OF NITROGEN CONTENT ON THE MECHANICAL PROPERTIES OF TiN x THIN FILMS PREPARED BY REACTIVE MAGNETRON SPUTTERING

Cyclic Stress-Strain Curve for Low Cycle Fatigue Design and Development of Small Specimen Technology

END FORMING OF THIN-WALLED TUBES USING A DIE

CHAPTER 9 INFLUENCE OF RESIDUAL STRESSES ON THE FAILURE PRESSURE OF CYLINDRICAL VESSELS

27th Risø International Symposium on Materials Science, Polymer Composite Materials for Wind Power Turbines, 2006

Part IA Paper 2: Structures and Materials MATERIALS Examples Paper 3 Stiffness-limited Design; Plastic Deformation and Properties

NUMERICAL MODELING OF BUCKLING OF THE LONGITUDINAL REINFORCEMENT IN BRIDGE COLUMNS

CH 6: Fatigue Failure Resulting from Variable Loading

Abstract. Introduction and Background

Fundamental Course in Mechanical Processing of Materials. Exercises

Transcription:

Supplementary Materials Hierarchical Nanolattice Design The hierarchical nanolattices fabricated in this work are designed using a recursive method that combines different unit cells into hierarchical geometries. The design process takes place as follows: 1) two (unique or identical) unit cell geometries are prescribed, 2) one unit cell is designated 1 st order and the other 2 nd order, and 3) the 1 st order unit cell is patterned along the length of the 2 nd order unit cell with N repeating units, resulting in a fractal-like geometry (Fig. S1). These steps can be repeated iteratively to create a fractal of any order, and the method is sufficiently general that it can be repeated for a wide range of unit cell geometries (Fig. S2). For the samples tested in this work, 1 st order axial support beams were added along the length of the 2 nd order beam to ensure that the hierarchical beams formed a stretching-dominated geometry. Figure S1: Hierarchical nanolattice design project showing the construction of an octahedron-of-octahedra unit cell. Figure S2: Hierarchical nanolattice geometries showing the versatility of the design process. The different geometries shown are: a) a cage of crosses, b) a cube of reinforced-bcc unit cells, c) an octet-truss of octahedra, and d) a cuboctahedron of embedded-octahedra. Fabrication Details Hierarchical nanolattices were fabricated out of solid polymer, polymer-ceramic core-shell composite, and hollow ceramic. There are three unique steps involved in the fabrication of these samples: 1) writing of a polymer sample, 2) coating of that polymer sample with a ceramic thin film, and 3) removal of the polymer. Each material system tested in this work represents one of the steps involved in the fabrication process. Fabrication of all samples begins with the writing of a polymer sample out of photoresist (IP-Dip 780) using two-photon lithography (TPL) direct laser writing (DLW) in a Photonic Professional DLW system (Nanoscribe GmbH). Samples are written using laser powers in a range from 6-12mW and a writing speed of ~50µm/s. The minimum writing spot size (voxel) is an ellipsoid with a width between 150-600nm and approximately a 4:1 height to width aspect ratio. The laser power is used to control the

diameter of the tubes, and the speed varies slightly during the writing process to control the quality of the structure. In the samples, individual 1 st order beams are written with three lines one central line and one on either side with a 150µm offset perpendicular to the axis of the beam to reduce the ellipticity of the beam. After a polymer scaffold is written, samples are conformally coated with 20nm of aluminum oxide (alumina) using atomic layer deposition (ALD). Deposition is done at 150 C in a Cambridge Nanotech S200 ALD system using the following steps: H 2 O is pulsed for 15ms, the system is purged for 20s, trimethyl aluminum (TMA) is pulsed for 15ms, the system is purged for 20s, and the process is repeated. The carrier gas is nitrogen, which is used at a flow rate of 20sccm (standard cubic centimeters per minute). The process was repeated for 200 cycles to obtain the desired thickness coating. The thickness of the coatings was verified using spectroscopic ellipsometry with an alpha-se Ellipsometer (J.A. Wollam Co., Inc.). After deposition, a focused ion beam (FIB) (Versa 3D DualBeam, FEI) is used to mill away selected regions of the sample to expose the polymer to air. Once the polymer is exposed, samples are placed into an O 2 plasma etcher (Zepto, Diener GmbH) for between 30-60 hours at a pressure of 0.5 mbar and at 100W of power in order to fully remove the polymer. It is possible to determine whether the polymer has been fully etched away by looking for any contrast change in the beams using a scanning electron microscope (1). Polymer Constituent Properties Mechanical characterization of the IP-Dip polymer was performed through micropillar compression experiments in a nanomechanical testing device (TI 950 Triboindenter, Hysitron Inc.). Micropillars were compressed using a 20µm diamond flat punch tip to 10-15% strain at a rate of 10-3 s -1 then held at their peak displacement for 50 s before unloading. Samples were fabricated out of IP-Dip photoresist using an identical DLW method to that described above for hierarchical nanolattices. Samples were fabricated and tested with diameters between 2-10µm, and length-to-diameter (L/D) ratios between 2 and 4. For each micropillar, stress-strain data was obtained and used to determine the Young s modulus (E) and compressive yield strength (σ y ). The Young s modulus is calculated using the slope of the linear regime of the stress-strain curve. The compressive yield strength is calculated by finding the intersection of the stress-strain data with a 0.2% strain offset curve from the linear regime. Three representative data sets along with their corresponding yield strength and stiffnesses are shown in Figure S3. The stress-strain data has an initial toe region followed by a linear regime and then a plastic flow region. The toe region is likely due to improper alignment or contact of the indenter tip with the sample, and was consequently ignored in the calculation of the Young s modulus. Based on the stressstrain data, we found that the IP-Dip polymer had an average modulus of E = 2. 1 ± 0. 3 GPa and an average yield strength of σ y = 67. 2 ± 4. 7 MPa (Fig. S3).

Figure S3: Representative stress-strain data for each slenderness ratio tested (L/D = 2, 3, 4) with various radii. The arithmetic mean of the Young s modulus (E) and compressive yield strength (σ y ) are plotted. The inset image shows a set of pre-compression micropillar samples (scale bar: 50 µm). Composite Constituent Properties The properties of the composite were calculated using a Voigt model rule of mixtures, with the properties of the ALD Al 2 O 3 taken from (Refs. 2 6). In the 2 nd order half cells, the polymer beams have dimensions of a = 753nm and b = 317nm, where a and b are the major and minor radii of the ellipse, respectively. The ceramic shell has a thickness of t = 20nm; the volume fraction of polymer in the beams can be calculated to be f = A polymer /(A polymer + A ceramic ) = 91.6%. Given the polymer properties of E p = 2.1 ± 0.3 GPa and σ yp = 67.2 ± 4.7 MPa, and the ceramic properties of E h = 165 GPa and σ yh = 5.2 GPa, we can obtain the core shell composite properties to be fe p + (1 f)e h = E c = 15. 8 GPa and fσ yp + (1 f)σ yh = σ yc = 509 MPa. Relative Density Calculations The relative density of each hierarchical sample was determined using a representative CAD model (SOLIDWORKS, Dassault Systèmes). The model enabled quick calculations for solid and hollow samples. Figure S4 illustrates the design process of the CAD model.

Figure S4: Illustration of the design approach used to construct a CAD model. The above images show the construction of a 2 nd order octahedron-of-octets unit cell. (1) Unique nodes that make up the unit cell. (2) Nodes are minimally patterned to create a unit cell beam. (3-4) Assembled nodes are patterned to create the full geometry. (5) Boundaries are cut to form the unit cell. (6) Finished unit cell whose volume properties can be measured. Failure Mode Formulation Failure in hierarchical nanolattices primarily initiated in 1 st order beams. In polymer samples, failure occurs via beam (Euler) buckling, while in composite samples failure occurs via brittle fracture. The continuous and serrated flow behavior of the hollow nanolattices indicates the occurrence of Euler buckling and shell buckling, the combination of which gives rise to a ductile-like deformation with a near 100% recovery. Large drops in the stress generally correspond to large Euler buckling events; smaller drops are more likely to coincide with localized shell buckling events. The failure criterion for beam buckling and shell buckling in an elliptical beam with a semi-major axis a, semi-minor axis b, thickness t, and length L are 3π σ beam = π2 EI (kl) 2 A π2 E ( 4 ab2 t) L 2 (πt(a + b)) = 3π2 E ab 2 4L 2 a + b (S1) E σ shell = 3(1 ν 2 ) (t a ) (S2) Here, E is the Young s modulus and ν is the Poisson s ratio of the material. By equating the failure criterion for beam buckling and shell buckling, it is possible to find a transition between the two. Here, we quantify the transition using a critical dimensionless parameter SB as 3π 2 E ab 2 4L 2 a + b = E 3(1 ν 2 ) (t a ) (S3)

t(a + b)l2 SB = ( a 2 b 2 ) = 3π2 4 3(1 ν2 ) 12.2 (S4) cr If the critical buckling transition value for a beam is SB 12.2, shell buckling will be the dominant failure mechanism over beam buckling, and if SB > 12.2, beam buckling will be the dominant failure mechanism. In the hollow samples tested in this work, the SB values range from 14.8 to 33.3, meaning that all the samples tested here are in a beam buckling failure mode regime. Previous work by Meza et. al. reported that hollow Al 2 O 3 nanolattices exhibit ductile-like behavior with near complete recoverability when the wall-thickness-to-tube-radius ratio is below a critical transition value of ( t ) = σ fs 3(1 a cr E ν2 ) 0.03, where σ fs is the fracture strength of the constituent material (1). All samples tested in this work have ( t ) < 0.03 < a (t), meaning that shell buckling will take place a cr preferentially over yielding. This analytical formulation shows that in the 1 st order beams, beam buckling will take place preferentially over shell buckling, and after the occurrence of beam buckling, shell buckling will take place over fracture, allowing samples to deform in a ductile-like manner and accommodate large strains. This matches very closely to what is observed experimentally. Full Hierarchical Nanolattice Mechanical Behavior 2 nd order full hierarchical nanolattices were fabricated in a 3x3x3 array using octahedra-of-octahedra hierarchical unit cells. Samples were compressed using an in-situ nanoindenter at a strain rate of 10-3 s -1 to 50% strain in a manner nearly identical to the half-cell samples. Failure mechanisms in the 2 nd order full nanolattices closely matched those in the 2 nd order half-cells for each material type: polymer samples underwent ductile failure via beam buckling, composite samples underwent catastrophic brittle collapse via fracture of the constituent Al 2 O 3, and hollow samples underwent serrated ductile-like failure via a combination of beam buckling and shell buckling. The stiffness and strength of the hierarchical nanolattice samples closely matched that of the half-cell samples, and a full list of results can be found in Figure S5. The primary distinction between the failure of the full samples and the half-cell samples is the emergence of a layer-by-layer collapse deformation mechanism, which led to high local stresses in the beams and reduced the recoverability in the full nanolattices compared with the half-cell samples. 2 nd order polymer nanolattices recovered immediately to 65-85% of their original height, with additional viscoelastic recoveries of 5-20% after unloading. 2 nd order composite nanolattices had a catastrophic strain burst with failure localized to the top of the lattice and exhibited little to no recovery. 2 nd order hollow nanolattices had layer-by-layer deformation that limited their recovery to only 60-75% of the original height, compared with 85-98% in the equivalent hollow half-cell samples. In hollow samples, the layer-by-layer collapse generated local highly strained regions where the buckled beams partially fractured without complete failure. The bifurcated beams experienced marginal recovery due to the residual strain energy being insufficient to return the sample to its original

configuration. Preventing layer-by-layer collapse would improve recovery but requires a more efficient hierarchical geometry to optimize load distribution. Figure S5: Comparison of mechanical data of 2 nd order half-cells against full hierarchical nanolattices showing their close agreement for (a) Young s modulus vs. relative density and (b) yield strength vs. relative density. Effect of Slenderness on Deformation Samples with a range of structural parameters and slenderness ratios were tested in this work. We show a number of representative samples of different material systems demonstrating the deformation and failure modes of samples with high and low slenderness in Figure S6 below. Figure S6: Compression experiments performed on 2 nd order half-cells showing the samples in an undeformed configuration, at 50% strain, and unloaded, along with the corresponding load (mn) v displacement (µm) data. All high slenderness samples are octahedra-of-octahedra and all low slenderness samples are octahedra-of-octets. Samples a-c have L = 8μm and N = 10, and samples d-f have L = 12μm and N = 20. Samples A and D are pure polymer, B and E are polymer-ceramic core-shell composites, and C and F are hollow Al 2 O 3.

FEM Modeling Modeling hierarchical lattices is a computationally demanding problem because the problem size scales exponentially. To reduce problem size one can use either a simplistic method or an adaptive method. The first, simplistic method uses bar elements in the Truss Model. Bar elements are the simplest and computationally least expensive method of modeling lattices. Our simple refined model is significantly more complex but can adaptively scale with the level of hierarchy through repeated use of substructure generation and utilization. Two modeling approaches were used to predict lattice stiffness and stress distributions. The first approach uses an in-house finite element code with small-strain bar elements having only axial stiffness. A linear elastic material model is assumed to govern the bar response. Figure S7 shows an example truss structure with the half-cell bottom fully constrained and the vertical displacements imposed on the top nodes of the cell. Vertical reaction forces were calculated at the top nodes to determine the lattice stiffness. Figure S7: Truss model of half-cell octahedron of octahedrons with applied boundary conditions: red all translation DOF constrained, blue prescribed vertical displacement. In contrast to the first model approach, the second, referred to as the refined model, is a scalable solution to modeling hierarchical lattices by using degree-of-freedom (DOF) condensation. The refined model consists of two stages: substructure generation and substructure utilization. Substructure generation begins with identifying the geometrically unique nodes and beams (Fig. 4). For compatibility between nodes and beams generalized DOFs, referred to as compatibility DOF, are introduced at the center of each node- and beam-connecting edge or face and the latter are assumed to deform rigidly; e.g. the end faces of each beam deform rigidly so that their motion can be described solely by its compatibility DOFs (i.e., its generalized translational and rotational DOFs). A substructure condensation method is applied to each beam or node part such that the only remaining DOFs are the compatibility DOFs. Substructure utilization then extracts the effective stiffness matrices of all

substructure parts, assembles those accordingly into a hierarchical half-cell lattice, and couples the compatibility DOF of each substructure to the appropriate adjacent substructure (Fig. 4D and E). Substructure generation and utilization can be performed repeatedly on increasingly larger levels of hierarchy to reduce the computational cost of evaluating the final lattice stiffness. Also, the extraction of the effective beam and node response from full-scale finite element models of the latter can be performed a priori so there is no need to recover full-resolution simulations during the final simulation run. Two general trends are observed for solid and hollow lattices. First, solid lattice stiffness predicted by the simple truss model is generally less than that of the refined model. Secondly, hollow lattice stiffness predicted by the truss model is greater than that of the refined model. These differences arise because of the contributions of bending and node junctions to degrading the overall lattice stiffness in the refined model. Figure S8 summarizes computational and experimental stiffness data and compares the simulated stiffnesses obtained from each model for geometrically perfect lattices. The refined model more accurately predicts properties of solid polymer and hollow ceramic lattice stiffness by 5.0 and 1.4 times, respectively, relative to the truss predictions. In simulating composites nanolattices, the truss model is 1.4 times more accurate than the refined model. If one takes into account geometric imperfections of the nanolattices, such as the beam waviness, the predicted stiffness of both the refined model and the truss model would decrease such that the refined model is always more accurate than the truss model relative to experimental data.

Figure S8: Comparison of model approximations and experimental data of second order half-cell lattices consisting of (a) solid polymer, (b) solid polymer with ceramic coating, (c) hollow ceramic.

Complete List of Samples Tested Composition Geometry Unit Cell Size (µm) Fractal Number Relative Density 8 10 1.03E-2 8 15 4.49E-3 Octahedron-of-Octahedra 8 20 2.50E-3 12 10 4.80E-3 12 15 2.09E-3 Solid Polymer 12 20 1.16E-3 8 10 2.10E-2 8 15 9.46E-3 Octahedron-of-Octets 8 20 5.34E-3 12 10 9.82E-3 12 15 4.41E-3 12 20 2.49E-3 8 10 1.12E-2 8 15 4.88E-3 Octahedron-of-Octahedra 8 20 2.72E-3 12 10 5.23E-3 12 15 2.28E-3 Composite 12 20 1.27E-3 8 10 2.28E-2 8 15 1.03E-2 Octahedron-of-Octets 8 20 5.83E-3 12 10 1.07E-2 12 15 4.81E-3 12 20 2.73E-3 8 10 8.95E-4 8 15 3.90E-4 Octahedron-of-Octahedra 8 20 2.17E-4 12 10 4.30E-4 12 15 1.87E-4 Hollow Al 2O 3 12 20 1.04E-4 8 10 1.82E-3 8 15 8.36E-4 Octahedron-of-Octets 8 20 4.89E-4 12 10 8.78E-4 12 15 3.95E-4 12 20 2.41E-4 Table S1: Full list of fabricated 2 nd order half-cell geometries with corresponding relative densities Composition Unit Cell Size (µm) Fractal Number Relative Density Solid Polymer 8 5 2.32E-2 3 10 3.55E-3 8 5 1.83E-2 Composite 8 5 2.51E-2 3 10 2.15E-3 Hollow Al 2O 3 8 5 1.83E-3 3 10 4.73E-4 Table S2: Full list of fabricated 3 rd order octahedron half-cell geometries with corresponding relative densities Composition Unit Cell Size (µm) Fractal Number Relative Density 8 10 1.03E-2 Solid Polymer 6 15 4.95E-3 4 20 3.26E-3 Composite 8 10 1.12E-2 6 15 5.49E-3 8 10 8.95E-4 Hollow Al 2O 3 6 15 5.35E-4 4 20 4.71E-4 Table S3: Full list of fabricated 2 nd order octahedron full-lattice geometries with corresponding relative densities

References 1. Meza LR, Das S, Greer JR (2014) Strong, lightweight, and recoverable three-dimensional ceramic nanolattices. Science 345(6202):1322 1326. 2. Berdova M, et al. (2014) Mechanical assessment of suspended ALD thin films by bulge and shaftloading techniques. Acta Mater 66:370 377. 3. Ilic B, Krylov S, Craighead HG (2010) Young s modulus and density measurements of thin atomic layer deposited films using resonant nanomechanics. J Appl Phys 108:1 11. 4. Jen S-H, Bertrand JA, George SM (2011) Critical tensile and compressive strains for cracking of Al2O3 films grown by atomic layer deposition. J Appl Phys 109(8):084305. 5. Tripp MK, et al. (2006) The mechanical properties of atomic layer deposited alumina for use in micro- and nano-electromechanical systems. Sensors Actuators A Phys 130-131:419 429. 6. Bauer J, et al. (2015) Push-to-pull tensile testing of ultra-strong nanoscale ceramic polymer composites made by additive manufacturing. Extrem Mech Lett.