development of lightweight materials by meso- and

152
Scholars' Mine Scholars' Mine Doctoral Dissertations Student Theses and Dissertations Fall 2020 Development of lightweight materials by meso- and Development of lightweight materials by meso- and microstructure control microstructure control Myranda Shea Spratt Follow this and additional works at: https://scholarsmine.mst.edu/doctoral_dissertations Part of the Materials Science and Engineering Commons Department: Materials Science and Engineering Department: Materials Science and Engineering Recommended Citation Recommended Citation Spratt, Myranda Shea, "Development of lightweight materials by meso- and microstructure control" (2020). Doctoral Dissertations. 2955. https://scholarsmine.mst.edu/doctoral_dissertations/2955 This thesis is brought to you by Scholars' Mine, a service of the Missouri S&T Library and Learning Resources. This work is protected by U. S. Copyright Law. Unauthorized use including reproduction for redistribution requires the permission of the copyright holder. For more information, please contact [email protected].

Upload: others

Post on 09-Nov-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Development of lightweight materials by meso- and

Scholars' Mine Scholars' Mine

Doctoral Dissertations Student Theses and Dissertations

Fall 2020

Development of lightweight materials by meso- and Development of lightweight materials by meso- and

microstructure control microstructure control

Myranda Shea Spratt

Follow this and additional works at: https://scholarsmine.mst.edu/doctoral_dissertations

Part of the Materials Science and Engineering Commons

Department: Materials Science and Engineering Department: Materials Science and Engineering

Recommended Citation Recommended Citation Spratt, Myranda Shea, "Development of lightweight materials by meso- and microstructure control" (2020). Doctoral Dissertations. 2955. https://scholarsmine.mst.edu/doctoral_dissertations/2955

This thesis is brought to you by Scholars' Mine, a service of the Missouri S&T Library and Learning Resources. This work is protected by U. S. Copyright Law. Unauthorized use including reproduction for redistribution requires the permission of the copyright holder. For more information, please contact [email protected].

Page 2: Development of lightweight materials by meso- and

DEVELOPMENT OF LIGHTWEIGHT MATERIALS BY MESO- AND MICRO­

STRUCTURE CONTROL

by

MYRANDA SHEA SPRATT

A DISSERTATION

Presented to the Graduate Faculty of the

MISSOURI UNIVERSITY OF SCIENCE AND TECHNOLOGY

In Partial Fulfillment of the Requirements for the Degree

DOCTOR OF PHILOSOPHY

in

MATERIALS SCIENCE AND ENGINEERING

2020

Approved by:

Joseph W. Newkirk, Advisor K. Chandrashekhara

Jeremy Watts Yijia Gu

Ronald O’Malley

Page 3: Development of lightweight materials by meso- and

© 2020

Myranda Shea Spratt

All Rights Reserved

Page 4: Development of lightweight materials by meso- and

PUBLICATION DISSERTATION OPTION

iii

This dissertation consists of the following four articles, formatted in the style used

by the Missouri University of Science and Technology:

Paper I, found on pages 13-33, has been published in the Rapid Prototyping

Journal.

Paper II, found on pages 34-63, has been submitted to Metallurgical and

Materials Transactions A.

Paper III, found on pages 64-97, has been accepted for publication by S/N

Applied Sciences.

Paper IV, found on pages 98 -125, has been submitted to Metallurgical and

Materials Transactions B.

Page 5: Development of lightweight materials by meso- and

iv

ABSTRACT

In this work, two lightweight structures - lattice structures and metal matrix

syntactic foams (MMSF) - were studied. Honeycomb lattices were manufactured by

powder bed selective laser melting (SLM) from 304L stainless steel. The wall thicknesses

of these structures ranged from 0.2 to 0.5 mm. Surface roughness was the primary cause

of dimensional mismatch between the expected and as-built structures with an average

wall thickness increase of 0.12 mm. The strength of the honeycombs increased with

increasing wall thickness. A feature of the SLM microstructure, the melt pool boundary,

was also studied as a part of this work. 3D models of the melt pool boundary network

were developed. The goal of this study was to determine if the melt pool boundary

significantly affects elastic or plastic properties in SLM materials. The study found a

negative correlation between the concentration of the melt pool boundary network and

the yield strength. The melt pool boundaries also seem to contribute to anisotropic plastic

deformation. The copper matrix syntactic foams studied in this work were manufactured

using low-pressure binder injection molding. The binder was a water-agar-glycerin gel

optimized to achieve the highest sintered density for metal matrix syntactic foams.

Syntactic foams use hollow or porous particles as their filler material. The two fillers

used here were porous silica and bubble alumina, selected for their extreme differences in

size, material, and porosity distribution. Three samples for each filler with increasing

filler volume fraction were manufactured. The specific strength of the copper-porous

silica samples increased with increasing volume fraction. The copper-alumina bubble

sample greatly improved the impact resistance of the foam at low filler volume fraction.

Page 6: Development of lightweight materials by meso- and

v

ACKNOWLEDGMENTS

My work has been greatly helped by the support and assistance of many people,

without which I would not have finished this dissertation.

I would first like to thank my advisor, Dr. Joseph W. Newkirk, for the opportunity

to pursue this research at Missouri S&T. His expertise, guidance, and feedback have been

invaluable over the course of my dissertation.

I would like to acknowledge the Missouri University of Science and Technology

for the funding I received as a part of the Chancellor’s Distinguished Fellowship. The

financial support allowed me the freedom to research my passion and the opportunity to

meet people from both my own and other fields.

I would like to thank Dr. KC for the support and advice he gave over our

collaborative projects, which have enriched my experience during my graduate studies.

I would also like to thank fellow graduate students, both in my own McNutt Hall

and in Toomey Hall, for many years of discussion and spirited debate over our research

topics.

Finally, I would like to thank my family and friends for their unending

encouragement and faith. A special thanks to my dear husband, Cody, whose wise

counsel and encouragement kept me going.

Page 7: Development of lightweight materials by meso- and

Vi

TABLE OF CONTENTS

PUBLICATION DISSERTATION OPTION...................................................................iii

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

ACKNOWLEDGMENTS.................................................................................................. v

LIST OF ILLUSTRATIONS............................................................................................. ix

SECTION

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

1.1. INTRODUCTION AND RESEARCH OBJECTIVES...................................... 1

1.2. LATTICE STRUCTURES.................................................................................4

1.3. METAL MATRIX SYNTACTIC FOAM.......................................................... 8

1.4. ORGANIZATION OF DISSERTATION........................................................ 11

PAPER

I. BUILD ACCURACY AND COMPRESSION PROPERTIES OF ADDITIVELY MANUFACTURED 304L HONEYCOMB............................................................ 13

ABSTRACT................................................................................................................. 13

1. INTRODUCTION.................................................................................................... 14

2. METHODOLOGY................................................................................................... 16

3. RESULTS AND DISCUSSION..............................................................................20

4. CONCLUSIONS...................................................................................................... 30

REFERENCES ............................................................................................................ 32

II. EFFECT OF THE MELT POOL BOUNDARY NETWORK ON THE ANISOTROPIC MECHANICAL PROPERTIES OF SELECTIVE LASER MELTED 304L........................................................................................................ 34

Page

Page 8: Development of lightweight materials by meso- and

ABSTRACT.................................................................................................................34

1. INTRODUCTION.................................................................................................... 35

2. MODELING............................................................................................................. 37

3. PROCEDURE ......................................................................................................... 44

4. RESULTS AND DISCUSSION..............................................................................47

4.1. EXPERIMENTAL WORK.............................................................................. 47

4.2. MODEL WORK...............................................................................................49

5. CONCLUSIONS...................................................................................................... 58

REFERENCES ............................................................................................................ 60

III. OPTIMIZATION AND CHARACTERIZATION OF NOVEL INJECTIONMOLDING PROCESS FOR METAL MATRIX SYNTACTIC FOAMS...........64

ABSTRACT ................................................................................................................ 64

1. INTRODUCTION.................................................................................................... 65

2. MATERIALS AND METHODS............................................................................. 69

2.1. POWDER CHARACTERIZATION................................................................ 69

2.2. BINDER DEVELOPMENT............................................................................. 70

2.3. INJECTION MOLDING.................................................................................. 71

2.4. SAMPLE POST-PROCESSING AND CHARACTERIZATION................... 73

3. CALCULATION..................................................................................................... 74

4. RESULTS AND DISCUSSION.............................................................................. 77

4.1. POWDER CHARACTERIZATION................................................................ 77

4.2. BINDER OPTIMIZATION.............................................................................. 80

4.3. INJECTION MOLDING.................................................................................. 84

vii

4.4. BINDER BURNOUT AND SINTERING 87

Page 9: Development of lightweight materials by meso- and

4.5. MECHANICAL TESTING............................................................................ 91

5. CONCLUSIONS.................................................................................................... 91

CONFLICT OF INTEREST........................................................................................ 92

APPENDIX.................................................................................................................. 93

REFERENCES ............................................................................................................ 94

IV. INFLUENCE OF FILLER MATERIAL AND VOLUME FRACTION ONCOPPER MATRIX SYNTACTIC FOAMS.......................................................... 98

ABSTRACT.................................................................................................................98

1. INTRODUCTION.................................................................................................... 99

2. EXPERIMENTAL PROCEDURE......................................................................... 102

3. RESULTS AND DISCUSSION............................................................................ 106

3.1. POWDER CHARACTERIZATION.............................................................. 106

3.2. SAMPLE CHARACTERIZATION............................................................... 108

3.3. MECHANICAL TESTING............................................................................ 114

4. CONCLUSIONS.................................................................................................... 119

ACKNOWLEDGEMENTS........................................................................................121

REFERENCES........................................................................................................... 121

SECTION

2. CONCLUSIONS AND FUTURE WORK...........................................................126

BIBLIOGRAPHY ........................................................................................................... 130

viii

VITA 138

Page 10: Development of lightweight materials by meso- and

ix

LIST OF ILLUSTRATIONS

Figure 1.1: Three unit cell designs are shown...................................................................5

Figure 1.2: An optical micrograph of the as-built SLM microstructure of 304L stainlesssteel....................................................................................................................7

Figure 1.3: A SEM image of a copper and alumina bubble syntactic foam...................... 9

PAPER I

Figure 1: The CAD model for the H2 honeycomb specimens.........................................17

Figure 2: The surface of a honeycomb wall.....................................................................21

Figure 3: The width of a wall can be indirectly calculated from the number of melt pool tracks across it, the hatch spacing, and the average width of a melt pool..................................................................................................................23

Figure 4: Etched micrographs of honeycomb specimens................................................. 25

Figure 5: Columnar grains that cross the melt pools were observed.................................26

Figure 6: A schematic of the locations from which hardness measurements were takenis shown............................................................................................................ 26

Figure 7: An example of the stress- displacement curves generated by the compressiontesting................................................................................................................27

Figure 8: A buckled wall after compression testing. T..................................................... 28

Figure 9: Yield and ultimate stress and specific yield stress of the honeycomb areshown............................................................................................................... 31

Figure 10: The yield stress and ultimate stress of the solid material are shown..............31

PAPER II

SECTION Page

Figure 1: The stripes scan pattern showing the input parameters used by the RenishawAM250............................................................................................................. 39

Figure 2: Optical images of SLM 304L in two magnifications.......................................40

Page 11: Development of lightweight materials by meso- and

x

Figure 3: The model of a single melt pool compared to the actual solidified melt poolin the 304L ss...................................................................................................41

Figure 4: The model of a single layer with two colors to indicate the change in laserdirection as the melt pools were created..........................................................42

Figure 5: The scan pattern rotation is shown beside the model for that scan patternrotation..............................................................................................................42

Figure 6: The single melt pool mesh for the 0° hatch angle model, before and afterBooleans are applied........................................................................................43

Figure 7: The three RVE models after the boundaries were removed.............................43

Figure 8: The 67° hatch angle model rotated around the [100] direction by 0°, 45°,and 90°.............................................................................................................44

Figure 9: The four build orientations of the cylinders are shown....................................46

Figure 10: Anisotropy in the 0°-[100]-0° specimens after compression testing..............50

Figure 11: The 0°-[010]-45° specimens sheared as the other 0° hatch angle specimensdid, but the (001) anisotropy was not elliptical............................................... 51

Figure 12: A hardness map of the 0°-[100]-90° specimen after compression testing.....52

Figure 13: There were two types of lines produced when the MPBN concentration wasplotted over the length of the model - flat and pseudo-sinusoidal.................53

Figure 14: The yield strength vs concentration of the MPNB in the compressiondirection are correlated.................................................................................... 54

Figure 15: The edges of the cylinders rippled after compression testing........................ 55

Figure 16: A close view of the melt pool boundary and cellular structure before (A)and after (B) compression testing....................................................................56

Figure 17: A 0°-[100]-0° specimen after compression testing shown from the side andthe top of the cylinder......................................................................................57

Figure 18: A 0°-[010]-90° specimen after compression testing shown from the sideand the top of the cylinder............................................................................... 58

PAPER III

Figure 1: The particle size distribution by count and volume is shown..........................79

Figure 2: SEM images of each powder are shown...........................................................80

Page 12: Development of lightweight materials by meso- and

Xi

Figure 3: The response surfaces for the a) wet density, b) dry density, c) bulk sintereddensity, d) open porosity after sintering........................................................82

Figure 4: The percent mass loss in each binder-only injection molded specimen overtim e................................................................................................................85

Figure 5: The percentage of binder in each specimen versus specimen number.............88

Figure 6: The percentage of porous silica in each specimen versus specimen number, ... 88

Figure 7: A element map of a 0% P-S specimen with the carbon, copper, and oxygenelements identified...........................................................................................89

Figure 8: A micrograph of a 15% P-S specimen showing the dispersion of the poroussilica particles and the morphology of the particles after sintering...............90

Figure 9: The strength and density of several MMSF materials from this work and theliterature...........................................................................................................92

PAPER IV

Figure 1: The particle size distribution by count percent and cumulative count percent for each powder where a) is copper, b) is alumina bubble, and c) is porous silica...............................................................................................................107

Figure 2: SEM images of the three powders - a) copper, b) silica, c) and d) alumina -show the morphology and relative size of the powders.................................108

Figure 3: The cross-sections of AB powder particles were used to determine thedensity of the AB powder.............................................................................. 109

Figure 4: Some of the AB particles in the copper matrix filled with copper powder....112

Figure 5: Elemental mapping of a 100% copper specimen shows the distribution ofcopper and oxygen in the sample...................................................................114

Figure 6: EDS map of the elements present in a P-S and copper specimen; elementsare indicated on the figure............................................................................. 115

Figure 7: EDS map of the elements present in an AB and copper specimen; elementsare indicated on the figure............................................................................. 116

Figure 8: The specific modulus and 0.2% offset yield strength for each sample.......... 117

Figure 9: The specific modulus of the porous silica and alumina bubble samplescontrasted with a model for metal foams by Zhu et al [42]......................... 118

Figure 10: The absorbed energy of each sample...........................................................120

Page 13: Development of lightweight materials by meso- and

LIST OF TABLES

Table 1: The nominal and measured dimensions for the honeycomb samples................. 18

Table 2:The measured wall thickness and cell size values, adjusted to correct forparticles sintered to the walls.............................................................................. 21

Table 3: The calculated width of the honeycomb walls using the number of melt pooltracks, width of an average melt pool, and hatch spacing.................................. 23

PAPER II

Table 1: The test matrix for each sample, where the orientation listed shows the planenormal to the compression direction.................................................................. 45

Table 2: The yield strength, aspect ratio horizonal plane of each cylinder after testingfor each specimen are shown.............................................................................48

Table 3: The maximum, minimum, and difference between the pseudo-sinusoidal|peaks for the three directions of interest.......................................................... 59

PAPER III

Table 1: Details provided by the manufacturer on each metal and ceramic powder usedin this experiment............................................................................................... 70

Table 2: The composition of each slurry in the injection molding reservoir beforeinjection molding is listed.................................................................................. 72

Table 3: The general composition of each sample for the model calculation.................. 75

Table 4: The composition of each binder for the model calculation, all units in volumepercent................................................................................................................ 75

Table 5: The 10th, 50th, 90th percentiles and average particle size in microns for eachpowde................................................................................................................. 80

Table 6: The cubic model b-values calculated for each response and the predictedvalue based on the optimized composition........................................................ 83

Table 7: The error for each response and relevant variance calculations are tabulated. .. 84

xii

PAPER I Page

Page 14: Development of lightweight materials by meso- and

Table 8: The average volume percent of binder and porous silica in each sample afterinjection molding............................................................................................... 87

Table 9: The relative density of each sample after each stage of sample post­processing shows the approximate amount of porosity in each sample...........89

PAPER IV

Table 1: The solids loading, filler, and copper volume percent of each sample............103

Table 2: The density and relative density for each stage and sample are shown withtheir standard deviation.....................................................................................110

Table 3: The volume percent of binder in each sample was used to calculate a moreaccurate relative density for each sample......................................................... 111

xiii

Page 15: Development of lightweight materials by meso- and

SECTION

1. INTRODUCTION

1.1. INTRODUCTION AND RESEARCH OBJECTIVES

Porosity is often considered a defect in a material, but with careful engineering, it

can be turned into a strength. Generally, structural properties such as strength and

stiffness increase as material density increases [1]-[6]. Material design charts produced

by Ashby [7] plot this relationship for many materials. The goal of lightweight structural

material design is to create materials have higher structural property to density ratios

compared to other materials. By controlling porosity or other density-decreasing

additives, such materials can be designed. [1], [2], [7]-[10]. This work investigates two

different structure-controlled materials - lattice structures and metal matrix syntactic

foams.

Lattice structures are porous foams with engineered geometry. [1], [10].

Periodicity is beneficial for these structures in that the properties are more predictable. A

unit cell, a shape translated and repeated in 2 or 3 dimensions where each cell is on the

micro to meso length scale, is one of the simplest lattice structure designs. This design is

often inspired by crystal lattice structures and trusses [1], [11]. Lattice structure designs

also include graded cellular structures, unit cells with curved surfaces, auxetic unit cells,

and more [2], [10]-[13]. Lattice structures are typically manufactured using either

investment casting or an additive manufacturing (AM) method. Selective laser melting

(SLM), a powder bed AM process, is often used to make complex lattice structures that

Page 16: Development of lightweight materials by meso- and

are not otherwise possible to fabricate. The powder bed process is ideal for these

structures as the powder can support the lattice while it is built. The literature includes

much research on the process-microstructure-property relationship, but there are aspects

of that relationship that are poorly defined [14]—[16]. One aspect involves a

microstructure feature known as the melt pool boundary. The fish-scale shape of a

solidified melt pool is a well-known feature of the SLM process, and it appears in nearly

every as-built, etched micrograph of SLM materials. The early literature on the melt pool

boundary indicates that may be a failure point in the structure [16]—[19]. The effect of the

melt pool boundary may be magnified in the thin features of the SLM lattice structures.

SLM-manufactured lattice structures are also vulnerable to dimensional mismatch and

surface roughness, demanding rigorous optimization and analysis be performed on the

structures [20].

Metal matrix syntactic foams are three-phase particulate composites. The MMSF

consists of a metal matrix embedded with hollow or porous filler particles — usually

ceramic or glass. Porosity in these composites is controlled by the particle size of the

filler. The filler surrounds the porosity, shielding the metal matrix from the pores. The

filler also reinforces the matrix as in solid metal matrix composites [8]. These materials

have a wide range of potential applications through the primary research vector has been

lightweight structural applications [21]-[25]. Heat sinks, electrical and vibrational

shielding, ballistic armor, packaging, and other applications are other potential uses for

these materials [4], [5], [26], [27]. The design of MMSF materials is limited by sources

of porous or hollow filler material, material compatibility between the phases, and

manufacturing the composites [9], [28], [29]. Porosity-containing filler powders are not

2

Page 17: Development of lightweight materials by meso- and

3

available in a wide variety of sizes, materials, or densities. The ones that are available are

not compatible with every matrix material or process. For good material compatibility,

both the filler and the matrix must survive the melting or sintering process without

undesirable diffusion, corrosion, or interfacial reactions. There are three main processing

techniques used for MMSF manufacturing. These are stir casting, pressure infiltration,

and binder injection molding [5]. Injection molding is the method used in this work. In

this technique, the metal and filler powders are combined with a binding agent to form a

slurry. This slurry is then injected into a mold. Parts are densified by sintering. Slurry

segregation, matrix porosity, and material compatibility during sintering are the key

challenges to overcome with this processing technique.

The knowledge gaps targeted in this research included the process-microstructure-

property relationship in SLM lattice materials and injection molded metal matrix

syntactic foams. This work expanded the body of research involving structure-controlled

lightweight structural materials. Four papers were written that each address a specific

objective, as follows:

1. Identify key build quality control issues in SLM lattice structures and study

their effects on the properties of the lattice.

2. Study an aspect of SLM microstructure that has received little attention in the

past, the melt pool boundary, and determine the effect of the melt pool boundary on

anisotropic material properties.

3. Develop a water-based binder for use in a low-pressure injection molding

processing, specifically optimized to decrease matrix porosity in a metal matrix syntactic

foam.

Page 18: Development of lightweight materials by meso- and

4

4. Determine the effect of filler material and volume percent on the properties and

manufacturing of metal matrix syntactic foams.

1.2. LATTICE STRUCTURES

Lattice structures are designed by manipulating the meso-scale geometry of the

material. Many unit cell lattice structures are trusses derived from crystal lattice

structures. Honeycombs are common 2D structures based on the HCP structure [30].

Cubic truss structures, based on BCC or FCC structures, are common 3D lattice unit cells

[10], [11], [31]-[33]. Other designs include curved surfaces such as gyroids and non­

crystal structure inspired truss designs like the octet truss [2], [10]—[12], [34], [35].

Figure 1.1 shows a honeycomb, a BCC truss, and a gyroid unit cell. The unit cell design

can be optimized by grading the unit cell density if, for example, certain areas of a part

need to take more load than others or if successive failure of the part is desired. Density

grading can be done by changing the strut size or the unit cell size in the target area [36].

A critical part of the lattice structure design process is modeling the mechanical

properties. The mechanical properties of lattice structures are geometry-dominated [30],

[34], [37]. For example, honeycomb structures are stronger vertically than in-plane with

the honeycomb octagons, regardless of the material used [38]. The unit cell properties

can be modeled with mathematics, but modern tools - such as finite element analysis

(FEA) - are more common [1], [10], [20], [30], [31], [34], [35], [37], [38]. A

representative volume element can be used to model the structure, rather than an entire

part. Unit cell designs are particularly useful for research and development if a part is not

yet envisioned for a specific application or if the part shape is confidential.

Page 19: Development of lightweight materials by meso- and

5

Figure 1.1: Three unit cell designs are shown- A) a 2D honeycomb, B) a BCC truss, andC) a gyroid.

Lattice structures can be manufactured using investment casting or an additive

manufacturing technique [10], [39]. Some 2D lattice structures can be made using

extrusion as well [39], [40]. Investment casting of lattice structures is time-consuming

and costly. The molds used for this method are made by coating a sacrificial polymer in

ceramic slurry. After the polymer is removed, the mold is filled with a suitably fluid

metal [4]. This method is well-understood in the literature, but the strut size and

complexity are limited by the flowability of the molten metal. Sheet folding and welding

can be used to make some lattice structure designs. In this method, sheets of metal are

deformed into the desired shapes then welded together. Powder bed selective laser

melting (SLM) is an excellent method of manufacturing lattice structures on a small scale

[10], [14]-[16]. The powder bed supports the lattice as it is built, and it can also be easily

removed from an open-celled lattice structure after building. The laser used as a heat

source is precisely controlled and capable of printing thin features on the order of a

hundred microns. The laser melts powder in a pre-determined path, layer by layer, with

fresh powder spread over the bed after each layer. This is known as the laser scan pattern.

Page 20: Development of lightweight materials by meso- and

The melt pool that forms from the melted powder solidifies rapidly as it cools. The size

and shape of the melt pool depends on the laser energy, residual heat from the

surrounding powder and dense part, and the properties of the material.

Research into the process-microstructure-property relationships in SLM parts is

active, and the majority is focused on the process parameter influence on properties [15],

[16], [18], [41]-[46]. The distinctive fish-scale microstructure of SLM materials is a

result of the laser scan pattern used to make the part [15], [16], [44], [45], [47], [48]. In

steel alloys, the solidified melt pools are outlined by a thin feature, called the melt pool

boundary, with a cellular sub-grain structure inside the melt pool. [16]—[19]. Figure 1.2

shows these microstructural features. Literature on the melt pool boundary seems to

indicate that it is a weak feature in the material. For example, cracks in fractured SLM

material may follow the melt pool boundary [19], [49]. Grains in SLM materials are often

columnar and grow in the build direction, often through several melt pools. These grains

are often oriented in the same direction [15], [16], [18], [48]. Porosity defects are usually

related to the powder, assuming other variables are optimized [45], [50], [51]. Gases and

water trapped between or inside particles can cause inclusions and gas porosity [11],

[51]-[53] All of these and more variables affect the final SLM material [47].

For lattice structures, the SLM process parameters are tuned for different goals

than in bulk parts [11], [54]-[56] Lattice structures have both high surface area and

geometry-dominated properties. The SLM process parameters should be tuned to reduce

dimensional mismatch between the intended and as-built geometry in these materials.

Dimensional mismatch can occur in several ways. Surface roughness can increase the

volume and mass of the structure enough to change the unit cell density. It also causes

6

Page 21: Development of lightweight materials by meso- and

7

Figure 1.2: An optical micrograph of the as-built SLM microstructure of 304L stainless steel, etched electrolytically in a 60:40 nitric acid: water solution.

local variation is strut width [20], [35]. The surface roughness of SLM parts comes in two

forms. First, un-melted powder welded to the surfaces of the struts is a relatively uniform

addition to the surface of the part. Second, melt pools at the edges of the part can also add

a rippled texture to the surface of the part from surface tension. Models that include a

variable strut diameter more accurately match experimental data, indicating that the

variation has a significant effect on the properties [31]. Dimensional mismatch can also

be caused by misprinting due to design limitations. Overhangs and smooth curved

surfaces can fail or include stair-stepping. Each machine has a lower limit on part width.

The melt pool size may be on the order of the width of a feature, which may mean that

the overlap between melt pools should to be adjusted to hit the target dimension of that

feature or vice versa [55]. Closed internal spaces will be filled with powder. For the

Page 22: Development of lightweight materials by meso- and

lattice structure, any unit cell should be tested for printability before printing a full part

[31], [33]. Also, it is often appropriate to attempt to model the actual as-built shape of the

printed part if time and cost allow [20], [55].

1.3. METAL MATRIX SYNTACTIC FOAM

Metal matrix syntactic foams (MMSFs) are particulate composites with

lightweight structural applications. The filler particles used in these materials contain

porosity which lightens the structure. Figure 1.3 shows an example of the MMSF

microstructure. MMSFs can be tailored to a specific application by changing the matrix

material, filler material, filler volume fraction, and filler shell thickness. The notable

lightweight structural properties of these materials are their excellent energy absorption,

specific strength, and specific stiffness [9], [24], [25], [28], [57]-[59]. These materials

can also have properties of both closed-celled foams and metal matrix composites as

well, such as wear resistance, vibrational damping, electrical and thermal conductivity,

and more [4], [26], [60]-[62].

Material compatibility between the filler and the matrix material is vital to the

quality production of MMSF parts. Many defects in these materials occur between the

matrix and filler during processing. Defects include unwanted porosity in the matrix,

fractured hollow spheres, unwanted matrix-filler interactions, and density gradients [21],

[25], [63]-[66]. The metal matrix is in contact with the filler during processing, usually at

high temperatures. This can cause chemical reactions between the matrix and filler -

inclusions, diffusion, and corrosion [21], [25], [28], [57], [67]-[69]. Certain metal-filler

combinations are particularly detrimental, such as alloys with very high oxygen affinities

8

Page 23: Development of lightweight materials by meso- and

9

Figure 1.3: A SEM image of a copper and alumina bubble syntactic foam.

(aluminum, for example) paired with oxide-containing ceramic or glass fillers. Beneficial

grain refinement can occur as well near the surface of the filler [26]. Because of their

relatively low softening temperature, many glass hollow spheres soften at the required

processing temperatures, which can collapse the internal porosity. [70], [71]. Pure silica,

which has a high softening temperature, can dissolve alloying elements from the metal

that lower the softening point of the glass [72].

There are two main types of porosity-containing filler materials available. First,

there are naturally occurring or by-product fillers. These include expanded clay/perlite

and fly ash cenospheres [29], [59], [73], [74]. The fly ash cenospheres in particular are

widely utilized MMSF filler materials because they are otherwise a waste product [9],

[75]. Naturally occurring fillers usually contain impurities; are often an aluminosilicate

with glassy phases; and may contain fractured particles and porous shell walls [22], [23],

Page 24: Development of lightweight materials by meso- and

10

[26], [28], [57], [60], [61], [65], [69], [76], [77]. Second, hollow filler materials

manufactured in a laboratory are available. Ceramic hollow spheres tend to be larger than

0.1 mm in size which may not be acceptable for all processing techniques [21], [25], [67],

[68]. Glass and metal hollow spheres are also available [58], [66], [70]-[72], [78]. Glass

filler can be available in smaller particle sizes than the ceramic spheres. The available

selection of hollow or porous filler limits the matrix materials and processes that can be

used [79].

The manufacturing of MMSF materials requires balancing several competing

factors. First, the filler particulates should be well-dispersed in the material. The filler

particles have very low density (less than 1 g/cm3) compared the matrix, so they tend to

float or segregate out of mixtures. Second, the particulates should keep their internal

porosity intact. The porosity can collapse if the particles soften or melt during processing.

The particles can fracture from mechanical stresses during processing. The matrix

material can infiltrate the pores if the shell wall is porous or corroded away. As a

corollary to the second factor: the selected manufacturing technique should be a near-net

shape process as subtractive machining will expose the filler porosity. Third, matrix

porosity should be eliminated from the material so that only the (controlled) filler

porosity remains.

With these criteria and restrictions, there are few processing techniques that can

be used to manufacture MMSF materials. Stir casting is one of the more popular methods

of making MMSFs in the literature [26], [57], [60], [61]. In this method, an impellor is

used to mix the filler into molten metal. The filler tends to float due to low density, so the

casting and cooling of the MMSF must be rapid. The impellor can fracture the filler due

Page 25: Development of lightweight materials by meso- and

to high shear forces. Further, this process only works with castable metals. Another

common processing method is infiltration [21]-[23], [25], [26], [28], [67], [68], [74],

[76], [80]. In this method, a bed of the filler is infiltrated by the molten metal with or

without pressure assistance. Infiltration can leave matrix porosity behind if the wetting

between the filler and the metal is poor or if the metal has a high viscosity. If the

infiltration pressure is too high, the filler can fracture. Powder metallurgy methods such

as binder injection molding are less common in the literature [78], [81]-[83]. In injection

molding, the filler, metal, and binder will tend to segregation while the slurry is non-solid

[66], [84]. Further, the hollow spheres can fracture if the shear forces during

compounding are too high [78]. The sintering temperature can also be an issue. If a

liquid-phase sintering process is used, it can affect the filler similar to the melt processes.

If no liquid phase is used and no pressure, then matrix porosity can be an issue.

1.4. ORGANIZATION OF DISSERTATION

An outline of this dissertation’s research objectives was given in Section 1.1. The

background research on lattice structures and metal matrix syntactic foams was reviewed

in Section 1.2 and 1.3. Four papers, each addressing one of the research objectives, were

included in this dissertation.

The first paper focused on the characterization and compression testing of SLM

honeycomb structures with increasing wall thickness. The dimensional mismatch

between the 3D model of the honeycomb structures and the as-built specimens was

thoroughly characterized. The effect of wall thickness on the mechanical properties of the

honeycomb lattice structures was investigated.

11

Page 26: Development of lightweight materials by meso- and

12

In the second paper, the melt pool boundary network, a microstructural feature of

the SLM process, was investigated. This feature is not well understood in the literature,

though a growing body of research indicates that it does influence the properties of SLM

parts. The melt pool boundary network was modeled using 3D software. The correlation

between this feature and the elastic and plastic properties of selective laser melted 304L

steel was investigated.

The third paper shifted the focus from lattice structures to metal matrix syntactic

foams. A water-based binder suitable for low-pressure injection molding of MMSF

material was developed. MMSF samples with a porous silica filler and a copper matrix

were produced and characterized.

The fourth paper used the process developed in the third paper to test the impact

of two different fillers on the foam properties. The porous silica and alumina bubble

powders were selected for their differences in size, material, and porosity distribution.

The contrast strongly showed the strengths and weaknesses of the two types of filler. The

effect of filler material and volume fraction on the yield strength and impact energy of

the MMSF materials was studied.

Page 27: Development of lightweight materials by meso- and

13

PAPER

I. BUILD ACCURACY AND COMPRESSION PROPERTIES OF ADDITIVELY MANUFACTURED 304L HONEYCOMB

ABSTRACT

Purpose: Thin walled metal honeycomb structures made with a selective laser

melting process were analyzed for their build quality and compression properties in this

study. The honeycombs were tested over increasing densities. Methodology: The density

of the honeycombs was changed by increasing wall thickness of each sample. The

honeycombs were tested under compression. Differences between the CAD model and

the as-built structure were quantified by measuring physical dimensions. The

microstructure was evaluated by optical microscopy, density measurements, and

microhardness. Findings: The Vickers Hardness (HV) of the honeycomb structures was

209±14 at 50g load. The compression ultimate and yield strength of the honeycomb

material were shown to increase as the wall thickness of the honeycomb samples

increased. The specific ultimate strength also increased with wall thickness, while the

specific yield stress of the honeycomb remained stable at 42 ± 2.7 MPa/g/cm3. The

specific ultimate strength minimized near 0.45 mm wall thickness at 82 ± 5 MPa/g/cm3,

and increased to 134 ± 3 MPa/g/cm3 at 0.6 mm wall thickness. Value: This work

highlights a single lightweight metal structure, the honeycomb, built by additive

manufacturing. The honeycomb is an interesting structure because it is a well-known

building material in the lightweight structural composites field, but is still considered a

Page 28: Development of lightweight materials by meso- and

relatively complex geometric shape to fabricate. As shown here, AM techniques can be

used to make complex geometric shapes with strong materials to increase the design

flexibility of the lightweight structural component industry.

1. INTRODUCTION

14

Additive manufacturing (AM) research continues to be popular as interest in

metal AM grows. However, the expensive trial and error adjustment of process

parameters to tune properties prohibits many industries from utilizing AM. [1], [2] Trial

and error procedures are usually done when there is not enough information about a

process to accurately model it. Accurate models require investigation into how process

parameters affect the as-built geometry and microstructure of a part and how those

features, in turn, affect the properties.[3] As accurate models of the AM processes are

built and validated, the cost to design and fabricate new components is reduced. This

paper explores how the as-built geometry of a lattice structure differs from the CAD

model, and how changing the geometry of the part affected the compression properties.

This structure was chosen due to increasing need for lighter transportation equipment -

especially in aerospace - which requires high specific stiffness and specific strength

materials. AM-built metal lattice structures are of particular interest due to the design

flexibility of AM.

Additive manufacturing is commonly used to make complex geometries -

sometimes with internal features or convoluted patterns. A gyroid is a good example of

one such shape. [4] One issue with making these parts is that many of them are not

Page 29: Development of lightweight materials by meso- and

machine-able after building. While this may be less of a problem in wire-fed systems,

many AM processes involve the use of metal powder. Powder particles can weld to the

surfaces of the part and become trapped in internal features during printing. Both of these

issues can affect the properties and performance of the part. Further, AM is a layer-by­

layer process. The layers create a ‘stair-step’ pattern on the surface of a part, contributing

to the surface roughness. Another issue is anisotropy due to the layering of the melt pools

- which adds the build direction to the list of variables that alter the mechanical

properties. All of these features can be controlled by process parameters, and powder (or

wire) quality to some extent. [5], [6] Models of melt pool formation and the resulting

effect on the surface quality have been done with some success. [7], [8]

Beyond the surface features, the internal features such as the solidification

structure, the dislocation and porosity distribution in the part, and the residual stress must

also be identified. [9] produced an in-depth characterization of the AM stainless steel

microstructure. These internal features are clearly tied to the process parameters (such as

laser power, scan speed, scan pattern, etc.), but how they are correlated is a complex

subject. For example, [10] found that remelting the layer increased density and reduced

surface roughness, and they determined the microstructural change associated with this

process. [11] found that residual stress can be controlled partially through part

orientation, energy density, and part spacing.

There are several particular challenges involved in additively manufacturing

lattices. For example, [12] found that there is a mechanical property drop when

transitioning from a thick to thin walled structures. This drop can be altered by improving

process parameters for thin walled parts. Their work suggests that surface quality is

15

Page 30: Development of lightweight materials by meso- and

increasingly important as wall thickness decreases. A design challenge in AM lattice

structures is that they must be self-supporting. A schoen gyroid and the schwatz diamond

are both examples of self-supporting lattices used in the literature. [13] Some lattice

geometries must be built in a particular orientation, or avoid a particular orientation.

Honeycombs, for example, cannot be built with the hexagons perpendicular to the build

plate without overhang problems. During mechanical testing, the strain can be difficult to

measure due to the lack of flat surfaces to which an extensometer can be attached. One

method of accurate strain measurement differential image correlation (DIC). [14] These

and other issues can make design and development of AM lattices difficult.

This study involves investigation of a component with a complex, but relatively

well-understood geometry. Two-dimensional honeycomb samples were printed in a

selective laser melting process. 304L stainless steel powder was used as the printing

medium. The build quality and microstructure were evaluated using a variety of methods

including hardness, microstructure, and dimensional differences between the as-built and

CAD model. The samples were then tested under compression.

2. METHODOLOGY

16

2D honeycombs were built with a Renishaw AM 250 powder bed system. 304L

stainless steel powder was used - a detailed analysis of which was published by [15]. The

powder was sieved before use to remove any particles above 63 microns in diameter. The

Renishaw uses a 1070 nm NdYAG laser. The following build parameters were kept

constant throughout the experiment: laser power of 200W, spot size of 70 p,m, layer

Page 31: Development of lightweight materials by meso- and

thickness of 50 p,m, hatch spacing of 85 microns, point distance of 60 microns, scan

speed of 800 mm/s, and flowing argon gas. It should be noted also that the Renishaw

AM250 uses a point-to-point laser pattern, not a continuous laser pattern.

An example of the CAD model for a honeycomb specimen can be seen in Figure

1. The figure shows labels for the cell size (C), wall thickness (t), height (Z), length (L),

and width (W) of the specimen. Four different samples were printed, with 15 specimens

per sample. The cell size remained constant at 3.97 mm, and the wall thickness was

varied from 0.1 to 0.5 mm. All the relevant CAD model dimensions are shown in Table

1. The specimens with 0.1 mm wall thickness were visibly porous, and were not used for

further experimentation.

17

WLT

C

Z

Figure 1: The CAD model for the H2 honeycomb specimens, where the relevantdimensions are labeled.

Calipers were used to measure the dimensions listed in Table 1. Each

measurement was taken three times in random positions across each of the 15 specimens

per sample set. Comparison between the as-built geometry and CAD model was done

using t-tests at 0.05 significance level. Two measurements of density were done. First,

Page 32: Development of lightweight materials by meso- and

Archimedes’ method was used, with isopropyl alcohol as the immersion liquid. Second,

the density of the honeycomb was found via a geometric measurement. This calculation

assumes the volume of the honeycomb as a cube.

18

Table 1: The nominal and measured dimensions for the honeycomb samples. Nominaldimensions are in bold.

H1 H2 H3 H4

Nominal Wall Thickness (mm) 0.2 0.3 0.4 0.5

Measured Wall Thickness 0 .3 ± 0 .0 1 0 .4 4 ± 0 .0 2 0 .5 2 ± 0 .0 2 0 .5 9 ± 0 .0 2

Nominal Cell Size (mm) 3.97 3.97 3.97 3.97

Measured Cell Size 3 .8 5 ± 0 .0 2 3 .8 2 ± 0 .0 3 3 .8 4 ± 0 .0 2 3 .8 3 ± 0 .0 2

Nominal Height (Z) 25.4 25.4 25.4 25.4

Measured Z (mm) 2 6 .7 3 ± 0 .0 9 2 6 .3 1 ± 0 .7 2 6 .8 6 ± 0 .0 7 2 6 .1 8 ± 0 .6 9

Nominal Width (W) 25.15 25.73 24.38 20.83

Measured W (mm) 2 5 .2 9 ± 0 .0 4 2 5 .8 4 ± 0 .1 5 2 4 .2 7 ± 0 .0 4 2 0 .9 9 ± 0 .0 4

Nominal Length (L) 26.67 27.28 26.67 22.86

Measured L (mm) 2 6 .6 5 ± 0 .0 5 2 7 .3 7 ± 0 .0 6 2 6 .7 3 ± 0 .0 5 2 2 .9 4 ± 0 .1 4

Microhardness was measured with ASTM E384-17. A Vicker’s indenter at 50g

load was applied to polished and etched specimens. The microhardness was found across

several specimens in multiple locations. Optical microstructure images of select

specimens were taken before and after compression testing. Specimens were

electrolytically etched in a 60:40 nitric acid: water solution.

Compression in the z-direction was done according to ASTM C365. Five

specimens from each sample set were tested. Specimens were loaded between two

parallel platens at a rate of 3 mm/minute. Strain was not directly measured, due to sample

Page 33: Development of lightweight materials by meso- and

19

constraints, but was instead measured from the LVDT of the machine and normalized to

the starting height of the specimen. Because of this, accurate measurement of the

modulus could not be done using this data. The compression stress was calculated using

Equation 1, noted as the stress of the honeycomb, and also by using Equation 2, noted as

the stress of the solid material. The specific stress was calculated using Equation 3. The

ultimate stress is defined as the maximum stress the honeycomb AM produced structure

withstood prior to plastic failure. The 2% offset yield stress was calculated using

MATLAB code.

FOi j —M W*L Equation 1

_ F*psas — W*L*pp Equation 2

X — P̂H Equation 3

where oh is the honeycomb stress, F is the load applied, Wis the width of the specimen, L

is the length, is the os is the stress in the cell walls, ps is the density of the solid material,

Ph is the density of the honeycomb, and X is the specific stress.

As stated, Equation 2 determines the stress in the solid material, or the stress in

the cell walls. The derivation of Equation 2 is as follows:

pos — — Equation 4

where As is the cross-sectional area of the solid portion of the honeycomb. As can be

derived from the density of the material as shown in Equation 5.

mPs H*A Equation 5

Page 34: Development of lightweight materials by meso- and

20

where H is the height of the honeycomb and m is the mass. The density of the honeycomb

(Equation 6) was used to bring Equation 2 into a similar form as Equation 1. Equation 7

shows the final form of As.

pH = ——— Equation 6h*w*l n

.Ac = ———— Equation 7 ̂ W*L*ps n

3. RESULTS AND DISCUSSION

Table 1 shows the CAD model dimensions of the honeycombs (notated as

nominal) and the average and standard deviation of the as-built dimensions. All measured

values were found to be significantly different from the CAD model dimensions (t-test,

a=0.05. The overall width and length of the honeycombs were less than 1% different than

the nominal dimensions. The cell size, height, and wall thickness measured dimensions

were all larger than the nominal dimensions by more than 1%. The most likely reason for

the height mismatch is a post-processing error. An EDM was used to remove the

specimens from the build plate, and care was not taken to perfectly match the CAD

model’s height. For the cell size and wall thickness dimensions, the presence of unmelted

particles sintered to the surface of the walls is likely the majority of the discrepancy.

These particles can be seen with the naked eye. Figure 2 is an example of the surface of a

honeycomb wall, as shown by an optical microscope with the ability to capture depth.

Because the maximum particle size of the powder is 63 microns in diameter, the cell size

and wall thickness measurements can be adjusted to account for the particles. These

adjustments, shown in Table 2, significantly reduced the differences between the nominal

Page 35: Development of lightweight materials by meso- and

21

and as-built dimensions. This is especially true of the cell size, where the difference was

reduced below 1%.

Figure 2: The surface of a honeycomb wall, shown to highlight the presence of particlessintered to the walls.

Table 2:The measured wall thickness and cell size values, adjusted to correct for particlessintered to the walls

H1 H2 H3 H4

Nom inal W all Th ickness (mm) 0.2 0.3 0.4 0.5M easured W all Thickness 0.3±0.01 0.44±0.02 0.52±0.02 0.59±0.02% Difference from Nom inal 51.0% 46.7% 29.8% 18.2%

Particle Size Adjustm ent 0.18±0.01 0.31±0.02 0.39±0.02 0.46±0.02

% Difference from Nom inal -12.0% 4.7% -1.7% -7.0%Nom inal Cell Size (mm) 3.97 3.97 3.97 3.97M easured Cell Size 3.85±0.02 3.82±0.03 3.84±0.02 3.83±0.02% Difference from Nom inal 3.1% 3.8% 3.3% 3.4%

Particle Size Adjustm ent 3.97±0.02 3.95±0.03 3.96±0.02 3.96±0.02

% Difference from Nom inal 0.1% -0.6% -0.2% -0.2%

Page 36: Development of lightweight materials by meso- and

22

There are potentially two other mechanisms that could contribute to the mismatch

between the as-built and nominal wall thickness. First, the natural surface roughness of

the wall, not including particles sintered to the wall. An accurate accounting of this

surface roughness could not be done due to the many particles clinging to the walls.

Second, the melt pool size and hatch spacing of the process are set prior to printing, and

were not adjusted to account for the thin wall specimens. It is likely that these melt pool

widths did not perfectly match the sizes intended by the CAD model. To check this, the

width of the honeycomb wall was calculated from the maximum width of the melt pool,

the number of laser tracks in each wall thickness, and the hatch spacing of the laser

tracks. The number of tracks is known from the microstructure, and the hatch spacing is a

set build parameter. The maximum width of the melt pool, however, is not easily

determined. The ‘top’ surface of the melt pool is the area of greatest width, and this area

is remelted in subsequent layers. Thus, direct measurement of the maximum width of the

melt pool is not possible using microstructure images. An approximation of the

maximum melt pool width can be made by measuring the width of the visible melt pool

structure. This was done with imageJ software and resulted in an approximate melt pool

width of 0.126±0.020 mm. The width of the wall can be calculated as shown in Figure 3

The melt pools are offset by the hatch spacing between laser tracks from the midpoint of

the melt pool. Table 3 shows the results of the calculation. The calculated widths are

within a standard deviation of the nominal, with the exception of the 5 mm specimen.

This data does indicate that for thin walled specimens, the wall thickness of the as-built

structure can differ from the CAD model simply due to the melt pool track width. Out of

hatch spacing, melt pool size, and number of laser tracks, the hatch spacing could be

Page 37: Development of lightweight materials by meso- and

altered to accommodate the thin walls the most easily. Any change to the build

parameters could significantly alter the density of the part, however, so care should be

taken.

23

Half melt pool width

Figure 3: The width of a wall can be indirectly calculated from the number of melt pool tracks across it, the hatch spacing, and the average width of a melt pool as shown in the

figure.

Table 3: The calculated width of the honeycomb walls using the number of melt pool tracks, width of an average melt pool, and hatch spacing.

H1 H2 H3 H4

Number of tracks 3 4 5 7

Nominal width (mm) 0.2 0.3 0.4 0.5

Calculated width (mm) 0.21 ± 0.2 0.30 ± 0.2 0.38 ± 0.2 0.55 ± 0.2

Page 38: Development of lightweight materials by meso- and

The density of the cell wall material - measured by Archimedes’ method - was

between 95% and 99% dense across all specimens. There was some evidence of lack of

fusion porosity in the walls. The density of the honeycomb, including the void space

within the cells, is listed in Table III. As the wall thickness increases, the density also

increases.

Images of the microstructure of the honeycombs can be seen in Figure 4. The

white arrow indicates the build direction and the black arrows indicate melt pool

boundaries. The dark phase visible in the etched micrographs is a cellular microstructure,

a result of fast solidification. It should be noted that two laser parameters were used to

build these structures. The first is optimized to reduce surface roughness, known as the

‘border scan’. This is done only on the outer surface of the component. The second set of

parameters is optimized to improve density and is done inside the component. Due to the

feature size of these parts, and the width of the melt pools, the border scan dominates the

majority of these structures. This could be detrimental to the properties of the honeycomb

samples. Columnar grains that crossed melt pools parallel to the build direction were

also observed as shown in Figure 5 shows some of these features.

The microhardness was evaluated for samples H1 and H2 in four different

locations- the border scan along the L-Z plane, the midpoint between two walls in the W-

L plane, the midpoint in the L-Z plane, and across the L-Z plane. A schematic of each

location is shown in Figure 6. The location of measurement did not appear to have a

significant effect on the test results (F-test, 95% confidence interval). The most notable

aspect of this test is that any difference between the border scan and the internal portion

of the cell wall could not be resolved using this test. This could indicate that the two

24

Page 39: Development of lightweight materials by meso- and

25

different scans did not notably affect the properties of the honeycombs. The 304L

honeycomb had a Vicker’s hardness (HV) of 209±14 at 50g load. Using the same set-up,

wrought 304 was also measured, with a HV of 237±9 at 50 g load.

BA

DC

20 pm

Figure 4: Etched micrographs of honeycomb specimens where build direction is indicated with white arrow and melt pool edges are pointed out with black arrows. A and B) 0.3

mm nominal thickness, image of the triple point where three struts meet. C and D) Crossection of a 0.4 mm nominal thickness honeycomb.

An example of the honeycomb stress-strain curves is shown in Figure 7. There are

six regions of interest indicated on the graph. Region 1 is the toe region, which is an

artifact of the compression testing equipment. Region 2 is the elastic region, from which

the modulus values were found. Region 3 is the yield stress of the honeycomb. Region 4

Page 40: Development of lightweight materials by meso- and

26

Figure 5: Columnar grains that cross the melt pools were observed, and two of these features are marked by black lines. The build direction is indicated by a white arrow.

Figure 6: A schematic of the locations from which hardness measurements were taken is shown. A) is the top of three honeycomb cells. B) is the cell wall of one honeycomb cell.

Straight black lines indicate where the hardness was measured.

Page 41: Development of lightweight materials by meso- and

27

is a plastic deformation region. Region 5 is the ultimate strength, and region 6 is the

typical cycling behavior seen in most honeycomb compression testing. Further work in

this area will involve finite element modeling to predict the properties of other wall

thicknesses and geometries.

Figure 7: An example of the stress- displacement curves generated by the compression testing. Six regions of interest are labelled.

Figure 8 shows an etched optical micrograph of a specimen after compression

testing. The failure of the honeycomb began at the mid-point of the cell walls by

plastically buckling. There was no evidence of cracking in the failure of these. An

interesting feature of this microstructure occurs near the compression side of the buckled

wall. The solidified melt pools along this side appeared to have been compressed and

elongated. This could indicate that there is a microstructure impact on these structures,

which typically have geometry-dominated properties. Work done by Wang et. al. [9] has

Page 42: Development of lightweight materials by meso- and

28

shown there is evidence that a large number of dislocations exist within the melt pools of

a similar alloy after laser additive manufacturing, but little to no dislocations in the melt

pool boundaries. This microstructure behavior, along with the unique crystallographic

grain orientation and crystallographic texture present in SLM material may have

influenced the failure of the honeycombs. [16], [17] A thorough review of the

compression properties of 304L after SLM processing is future work for this project.

Figure 8: A buckled wall after compression testing. The build direction is indicated by anarrow.

Figure 9 includes the stress data for the honeycomb samples, using Equation 1 to

calculate the stress in the sample. R-squared values on the figure indicate the fit to either

a 2nd order polynomial line or a linear line. The yield stress in the honeycomb appears to

increase linearly with wall thickness. The specific yield strength appears to stay relatively

level in the range tested. This indicates that the yield is noticeably influenced by the

Page 43: Development of lightweight materials by meso- and

29

geometry of the specimen, but that the density increase with increasing wall thickness is

approximately equal to the yield increase. The ultimate strength of the honeycomb

appears to increase quadratically with the wall thickness. The specific ultimate strength

appears to do the same. The specific strength appears to reach a minimum between H1

and H2, but further testing is needed to confirm this. It is unknown, but expected, that a

maximum specific strength will occur outside the tested range. This would indicate an at

least 3rd order dependence on the wall thickness. More testing outside the range tested

here is needed to confirm this suspicion. Regardless of the trends, the specific properties

remained higher than fully dense 304L stainless steel, which has a specific yield of 26

MPa/g/cm3 and a specific ultimate strength of 70 MPa/ g/cm3 (both values in tension)

according to the ASM aerospace specification. [18]

Figure 10 includes the stress data for the honeycomb samples, using Equation 2

to calculate the stress. This gives the stress experienced by the cell walls during testing.

The yield stress within the cell walls remains relatively stable across the specimens. This

is indicating the yield stress of the AM 304L, which should remain constant. The tensile

yield strength of 304L with 0% cold working is 210 MPa. [18] The yield strength of the

AM material is higher than the conventional material at around 320 MPa, possibly due to

the large number of dislocations that typically form in AM stainless steels. [9] The

ultimate strength in the cell walls increases quadratically as the wall thickness increases,

however there appears to be a minimum between H1 and H2, similar to the specific

strength of the honeycomb. This data indicates that there is a relationship between the

geometry of the specimen and the ultimate strength. The failure and failure prediction of

these honeycombs was investigated in depth by Anandan et. al. [19].

Page 44: Development of lightweight materials by meso- and

30

4. CONCLUSIONS

The thin-walled honeycombs manufactured in this study showed evidence of

deviation from the intended structure. This deviation exists both in the overall dimensions

of several samples and in the wall thickness and cell size of all the samples. Much of this

deviation can be attributed to particles sintered to the outer surfaces of the honeycombs.

The microstructure of the honeycombs was a typical AM microstructure. The

compression strength of the honeycomb and the solid material both increased from 92 to

267 MPa and 687 to 1052 MPa, respectively, with increasing wall thickness. The specific

compression strength of the honeycombs appears to reach a minimum of 82 ± 5

MPa/g/cm3 at 0.44 mm wall thickness with an increase to 134 ± 3 MPa/g/cm3 at 0.6 mm

wall thickness. The yield stress of the cell wall material was stable at 319 ± 16 MPa and

the yield stress of the honeycomb structure increased from 45 MPa to 83 MPa over the

wall thicknesses tested. The specific yield stress of the honeycomb structure was stable at

42 ± 3 MPa/g/cm3. Microstructural evaluation of the compression tested samples

identified the failure as plastic buckling of the cell walls.

Page 45: Development of lightweight materials by meso- and

31

140

120

100

80

60

40

20

0

160

Figure 9: Yield and ultimate stress and specific yield stress of the honeycomb are shown (Equation 1 was used to calculate the stress).

c3

xfixfiCD

1200

1000

800

600

400

200

0

Yield Stress

• Ultimate Stress

: * ▼

0.9476 * ^

R2 = 0.3541 ■*----------------- ♦

T-----1-----1-----1-----1-----1-----1-----1-----1

0.3 0.35 0.4 0.45 0.5 0.55 0.6Measured Wall Thickness (mm)

Figure 10: The yield stress and ultimate stress of the solid material are shown (Equation 2was used to calculate stress.)

Spec

ific

Stre

ss (M

Pa/d

ensit

y)

Page 46: Development of lightweight materials by meso- and

32

REFERENCES

[1] P. Hanzl, M. Zetek, T. Baksa, and T. Kroupa, “The Influence of Processing Parameters on the Mechanical Properties of SLM Parts,” ProcediaEng., vol. 100, pp. 1405-1413,Jan. 2015.

[2] Y. Huang, M. C. Leu, J. Mazumder, and A. Donmez, “Additive Manufacturing: Current State, Future Potential, Gaps and Needs, and Recommendations,” J.Manuf. Sci. Eng., vol. 137, no. 1, p. 014001, Feb. 2015.

[3] M. M. Francois et al., “Modeling of additive manufacturing processes for metals: Challenges and opportunities,” Curr. Opin. Solid State Mater. Sci., vol. 21, no. 4, pp. 198-206, Aug. 2017.

[4] C. Yan, L. Hao, A. Hussein, and D. Raymont, “Evaluations of cellular lattice structures manufactured using selective laser melting,” Int. J. Mach. Tools Manuf., vol. 62, pp. 32-38, 2012.

[5] S. Karnati, I. Axelsen, F. F. Liou, and J. W. Newkirk, “Investigation of tensile propertie sof bulk and LSM fabricated 304L stainless steel using various gage length specimens,” in Solid Freeform Fabrication Proceedings, 2016, p. 13.

[6] M. Jamshidinia and R. Kovacevic, “The influence of heat accumulation on the surface roughness in powder-bed additive manufacturing,” Surf. Topogr. Metrol. Prop., vol. 3, no. 1, p. 014003, Feb. 2015.

[7] G. Strano, L. Hao, R. M. Everson, and K. E. Evans, “Surface roughness analysis, modelling and prediction in selective laser melting,” J. Mater. Process. Technol., vol. 213, no. 4, pp. 589-597, Apr. 2013.

[8] S. A. Khairallah, A. T. Anderson, A. Rubenchik, and W. E. King, “Laser powder- bed fusion additive manufacturing: Physics of complex melt flow and formation mechanisms of pores, spatter, and denudation zones,” Acta Mater., vol. 108, pp. 36-45, Apr. 2016.

[9] Y. M. Wang et al., “Additively manufactured hierarchical stainless steels with high strength and ductility,” Nat. Mater., vol. 17, no. 1, pp. 63-71, Oct. 2017.

[10] E. Yasa and J.-P. Kruth, “Microstructural investigation of Selective Laser Melting 316L stainless steel parts exposed to laser re-melting,” Procedia Eng., vol. 19, pp. 389-395, Jan. 2011.

Page 47: Development of lightweight materials by meso- and

33

[11] A. S. Wu, D. W. Brown, M. Kumar, G. F. Gallegos, and W. E. King, “An Experimental Investigation into Additive Manufacturing-Induced Residual Stresses in 316L Stainless Steel,” Metall. Mater. Trans. A, vol. 45, no. 13, pp. 6260-6270, Dec. 2014.

[12] B. Brown, W. Everhart, and J. Dinardo, “Characterization of bulk to thin wall mechanical response transition in powder bed AM,” RapidPrototyp. J., vol. 22, no. 5, pp. 801-809, Aug. 2016.

[13] A. Hussein, L. Hao, C. Yan, R. Everson, and P. Young, “Advanced lattice support structures for metal additive manufacturing,” J. Mater. Process. Tech., vol. 213, pp. 1019-1026, 2013.

[14] V. Br0tan, O. Fergani, K. S0rby, and T. Welo, “Mechanical Properties of Biocompatible 316L Steel Rhombic Dodecahedron Lattice Structures,” in Solid Freedom Fabrication, 2016, pp. 2087-2094.

[15] C. S. Kriewall, A. T. Sutton, M. C. Leu, J. W. Newkirk, and B. Brown, “Investigation of heat-affected 304L ss powder and its effect on built parts in selective laser melting,” in Solid Freeform Fabrication Symposium, 2016, pp. 625­639.

[16] Y. Kok et al., “Anisotropy and heterogeneity of microstructure and mechanical properties in metal additive manufacturing: A critical review,” Mater. Des., vol. 139, pp. 565-586, Feb. 2018.

[17] T. DebRoy et al., “Additive manufacturing of metallic components - Process, structure and properties,” Prog. Mater. Sci., vol. 92, pp. 112-224, Mar. 2018.

[18] “ASM Material Data Sheet.” [Online]. Available: http://asm.matweb.com/search/SpecificMaterial.asp?bassnum=MQ304L.[Accessed: 25-May-2018].

[19] S. Anandan et al., “Failure In metal honeycombs manufactured by selective laser melting of 304 L stainless steel under compression,” VirtualPhys. Prototyp., vol. 14, no. 2, pp. 114-122, Apr. 2018.

Page 48: Development of lightweight materials by meso- and

34

II. EFFECT OF THE MELT POOL BOUNDARY NETWORK ON THE ANISOTROPIC MECHANICAL PROPERTIES OF SELECTIVE LASER

MELTED 304L

Myranda Spratt1; Joseph W. Newkirk1; Okanmisope Fashanu2; K. Chandrashekara2

1Materials Science and Engineering Department, Missouri University of Science andTechnology, Rolla, MO 65409

2Mechanical Engineering Department, Missouri University of Science and Technology,Rolla, MO 65409

ABSTRACT

Anisotropic mechanical properties are a well-known issue in selective laser

melted materials. The microstructure produced by selective laser melting is directional,

including the solidified melt pool structures and the grains. In this work, the melt pool

boundary's effects on 304L stainless steel's compressive properties were investigated. The

melt pool boundary effects were investigated experimentally and numerically. 304L

stainless steel solid cylinders were built in four directions using three hatch angle

rotations - 0°, 67°, and 105°. The twelve samples were compression tested, and the

results were analyzed. Results showed that both the hatch angle and build orientation

influenced the concentration of melt pool boundaries present in the manufactured

samples. A negative correlation of compressive strength to the melt pool boundaries'

concentration was also observed, indicating that the melt pool boundary negatively

affected the material's strength. Local anisotropic plastic deformation was also observed

in the 0° hatch angle samples in that the aspect ratio of the compressed cylinders was

greater than one. In those samples, it was observed that directions that plastically

Page 49: Development of lightweight materials by meso- and

deformed more contained a higher concentration of the melt pool boundaries than those

that deformed less.

35

1. INTRODUCTION

Additive manufacturing (AM) is a useful method of fabricating components with

complex internal features. AM processes are particularly efficient in fabricating low-

volume, high-value components. [1]-[4] One popular AM technique for making metal

components is selective laser melting (SLM) [4]. In SLM, a powder bed is used to

support the structure as it is being built. The laser melts the powder in a specified pattern

which creates the part layer-by-layer. As the laser melts the powder, a melt pool is

created. The melt pools are overlapped to create the dense part, remelting some of the

previously melted material. This precisely controlled process can be used to create

intricate structures with thin walls, such as lattice structures, as well as bulk parts [5]-[9].

The microstructure of SLM materials has a distinctive fish-scale pattern [6], [10],

[11] . In steel alloys, the fish-scale pattern appears as a thin band around the edge of the

melt pool, known as the melt pool boundary (MPB). The interior of the melt pool often

solidifies into a cellular structure [6], [12]. The different features in the solidified melt

pool may or may not include a detectable chemical difference. Rapid cooling (recorded at

104 K/s for SLM [13]) leaves a large number of dislocations and residual stresses [6],

[12] , [14]. The combination of heat from the laser and the material’s thermal properties

adjusts the size and shape of a melt pool as it forms and the phases that form as it cools

[15], [16]. This melt pool is curved - a result of the Gaussian laser beam distribution and

Page 50: Development of lightweight materials by meso- and

surface tension [12], [15], [17]. The material solidifies from the edges of the melt pool

towards the interior of the melt pool. Often, grain will grow epitaxially in the build

direction through several melt pools [12], [13], [18]. These grains often have the same

crystallographic orientation [12].

The relationship between the SLM process, the microstructure, and the properties

is complex and subject to much investigation [6], [12], [13]. Anisotropy in particular has

been recorded in strength, fatigue, and other properties of SLM parts, across many alloys

[12], [13], [19]-[24]. The build orientation and the laser scan pattern are notable process

parameters that induce anisotropic properties. This is likely due to the influence these

process parameters have on the microstructure of SLM material [12], [13], [15], [25],

[26]. The laser scan pattern is the path the laser takes in each layer. In many cases, this

path is repeated in each layer, as much as the shape of the part allows, only rotated in

each subsequent layer to reduce periodicity. The build orientation of the part is the

orientation of the part on the build plate. These parameters determine the pattern the melt

pools form in the material. In their review of anisotropy in AM materials, Kok et al [19]

concluded there three microstructural factors that influence anisotropy in SLM materials.

These are crystallographic texture, lack-of-fusion porosity, and columnar grains. With the

possible exception of crystallographic texture, these features can be controlled with build

parameters. Lack-of-fusion can be reduced by changing hatch spacing and by increasing

melt pool depth. The grain size is controllable by varying laser power and scan speed

[19], [27].

There is a second school of thought regarding anisotropy which was explored by

Shifeng et al [10]. In their work, the melt pool boundary (MPB) was considered as the

36

Page 51: Development of lightweight materials by meso- and

37

major influence on anisotropy. In this consideration, the MPBs allow for preferential slip

during deformation due to their low dislocation density [28]. This theory is supported by

Mower and Long [29], who concluded that the low fatigue strength and planar fracture of

their 316L samples was a result of the weak ‘build plane’. This theory was used to

explain the fracture behavior seen by Shifeng et al [10] as the crack path appeared to

follow the MPBs, rather than the grain boundaries.

In this work, the plastic deformation behavior of solid 304L cylinders was

analyzed. This work was initially done to support the development of a honeycomb

lattice structure failure model, published by Anandan et al [30]. During testing, the plastic

deformation of the cylinders was seen to be highly and visibly anisotropic. Further

investigation of the phenomenon has led to the investigation of the influence of build

orientation and scan pattern rotation on the elastic and plastic mechanical properties of

304L steel. Fashanu et al [31] investigated this topic from a property - parameter point of

view. This work aims to correlate property - microstructure- and parameters. A model of

the melt pool boundary network was created to facilitate this understanding, and the

results of this model were correlated to experimental test results.

2. MODELING

The melt pool boundary network (MPBN) was modeled using Blender software.

This software was chosen due to the relatively simple to use interface, ability to create

and modify complex meshes, and movie-making abilities. In this software, the scale can

be set by the user. The camera can also be set to an orthographic or a perspective mode

Page 52: Development of lightweight materials by meso- and

depending on the needs of the user. In this case, orthographic mode was selected and a

metric scale was used. Orthographic mode ensures that the mesh’s real size is imaged by

the software, rather than the perspective size. The metric scale ensures that any

measurements are in a convertible scale.

The MPBN depends on the laser scan pattern that is being used. The Renishaw

AM 250 machine used in the experimental part of this study uses a pulsed rather than the

more common continuous laser. The laser turns on and ramps to the correct wattage,

holds in place for the specified exposure time, then turns off. It then moves to the next

position at a set speed and turns on again. This means that the melt pool that forms under

the laser is not necessarily continuous. The melt pool shape will change depending on the

scan pattern used by the laser. In this case, the scan pattern was stripes. Figure 1 shows an

example of the stripes scan pattern. This figure also shows the spacing between laser

points and the hatch spacing between rows in a single layer. The hatch spacing in this

experiment was 85 microns, the point distance was 60 microns, and the layer thickness

was 50 microns.

The orientation of the developed models were based on the Renishaw builds. The

scan pattern of the laser always started the build by moving in the x or [100] direction,

and the [001] direction is the build direction or vertical axis. Each model was built with

this orientation. When planes or directions are mentioned, these are considered global

axes tied to the build plate. So, the build direction is always the [001] axis, and the layers

are in-plane with the (001) plane. The shape of the melt pool used in the models after

solidification was based on etched images of a 304L stainless steel (ss) specimen shown

in Figure 2. The idea to model the melt pool using this approach came from Li et al [16].

38

Page 53: Development of lightweight materials by meso- and

They modeled the thermal history and size change of melt pools using different input

parameters on the same machine and material as used here. The model boundary was

given a thickness to simulate the actual melt pool boundary. The melt pool boundary is

the material that solidifies first at the edges of the melt pool. In this work, the thickness

was set a single width of 0.1 microns.

39

Figure 1: The stripes scan pattern showing the input parameters used by the RenishawAM250 as modeled in this work.

The representative melt pool used in this model does not consider variations in the

melt pool size and shape. Melt pool variations can occur if the local energy density

fluctuates, border scans are used, the laser direction changes, or some other variation in

Page 54: Development of lightweight materials by meso- and

the machine occurs. Melt pool variations can include inclusions formed by residual

oxygen or carbon and lack of fusion porosity. Figure 3 shows the microstructure of a

typical 304L part as manufactured by the Renishaw AM250. Key features of the

microstructure are noted in the figure.

40

\ /

Porosity

| PMelt pool K _ „ ,, , \ Cellular boundary ^ Structure

1 0 0 p m 2 0

Figure 2: Optical images of SLM 304L in two magnifications etched with a 60:40 water: nitric acid electrolytic etchant. Relevant features have been pointed out on the images.

The first step of making the model was creating the model of a single melt pool.

The second step is to create the first layer of the model. This is done by arranging single

melt pools in a single plane to mimic the laser scan pattern. Each subsequent melt pool

annihilates any overlapping area with previous melt pools. This simulates the re-melting

that occurs. Figure 4 shows a single layer and the scan pattern that was mimicked to

create that layer. The alternating colors indicate the alternating direction the laser was

moving when that row was created.

Page 55: Development of lightweight materials by meso- and

41

(1 0 0 ) 50 [jm (0 1 0 ) Mt-.1 - 50 jjm {

Figure 3: The model of a single melt pool compared to the actual solidified melt pool inthe 304L ss used in this work.

The third step was to create multiple layers. Typically, the scan pattern is rotated

between layers. The rotation increases the ‘randomness’ of the structure. In this work,

three models were developed with different rotation angles. These were 0° (or non­

rotating), 67°, and 105°. Figure 5 shows each scan pattern and the corresponding model.

Six layers were created for each model.

In steps 2 and 3, the melt pool remelting was simulated by annihilating

overlapping material. This was done by using Boolean modifiers in the Blender software.

The Boolean modifier modifies a mesh by another mesh. Figure 6 shows a mesh before

and after a Difference Boolean modifier is applied.

Once each model of the MPBN was created, it could then be analyzed. In this

work, the concentration of the MPBN was analyzed through the structure. This was done

by slicing each model in the (001) plane in increments along the [001] direction. Each

model was sliced 60 times at regular intervals. Before slicing, the model itself was cut

down into a representative volume element (RVE). This RVE contained only overlapping

Page 56: Development of lightweight materials by meso- and

42

Figure 4: The model of a single layer with two colors to indicate the change in laser direction as the melt pools were created.

67'0' 150°

§//=■

Figure 5: The scan pattern rotation is shown beside the model for that scan patternrotation.

Page 57: Development of lightweight materials by meso- and

43

Figure 6: The single melt pool mesh for the 0° hatch angle model, before and afterBooleans are applied.

melt pools and eliminated any melt pools that were along the edges of the model. Figure

7 shows each of the RVE models.

Figure 7: The three RVE models after the boundaries were removed.

The concentration of MPBN in each slice when the model is oriented with the

build direction parallel to the [001] axis only shows the concentration of MPBN along

that plane. Other orientations have other MPBN concentrations. To track this, the model

was tilted around the [100] and [010] directions 5° increments from 0° to 90°. Figure 8

shows examples of the tilted model. Each tilted model was sliced and imaged in the same

manner as the 0° model. The model, when rotated around the [100] direction by 90°, was

sliced along the (010) plane. When it was rotated around the [010] direction by 90°, it

Page 58: Development of lightweight materials by meso- and

44

was sliced along the (100) plane. Each image was fed into FIJI software. This software is

capable of measuring the color data in each pixel of an image. This was done to each of

the image slices. The background color of the images was separated from the MPBN

color of the images, and the total number of MPBN pixels was found for each image.

This number was then divided by the total number of pixels in the model at that slice to

get the concentration as a percentage.

Figure 8: The 67° hatch angle model rotated around the [100] direction by 0°, 45°, and 90° where the lines are example slices through the models.

3. PROCEDURE

The procedure used to make the samples was detailed in Fashanu et al’s [31]

work. Briefly, 304L cylinders were printed with a Renishaw AM250. The argon-gas

Page 59: Development of lightweight materials by meso- and

atomized powder was sourced from LPW. The scan pattern used was a border scan

followed by the stripes scan pattern. Octagonal prisms with a height of 27.8 mm and a

side length of 3.84 mm were printed in 4 build orientations and 3 hatch angles. Three

replications were done for each sample. The test matrix is shown in Table 1. Figure 9

shows the four build orientations as they appear on the build plate. The hatch angles were

0°, 67°, and 105° rotation between layers. The prisms were removed from the build plate

with wire electron discharge machining (EDM). They were then machined to a diameter

of 6.35 mm and a height of 6.35 mm.

The samples were characterized before testing. Hardness measurents were carried

out on one specimen from each sample set. All Vicker’s hardness measurements were

done at 9.81 N for 10 seconds. Vicker’s hardness were done in three positions near the

center of the machined top surface of the cylinders for each sample. Archimedes’ density

was performed on the samples using water as the liquid medium, and the vacuum method

to remove gas bubbles.

45

Table 1: The test matrix for each sample, where the orientation listed shows the plane normal to the compression direction. The notation for the orientation is hatch angle -

rotation direction - rotation degree.

Sample (100) (010) (001) (011)

0° 0°-[010]-90°

0°-[100]-90°

0°-[100]-0°

0°-[010]-45°

67° 67°-[010]-90°

67°-[100]-90°

67°-[100]-0°

67°-[010]-45°

O o 105°-[010]-90°

105°-[100]-90°

105°-[100]-0°

105°-[010]-45°

Page 60: Development of lightweight materials by meso- and

46

[001]

Figure 9: The four build orientations of the cylinders are shown.

Compression testing was done as described in ASTM E3 using self-aligning

platens. The samples were loaded at a rate of 5x10"3 mm per min. The cylinders were

compressed to well past the yield point. A Visual Basic for Applications (VBA) code was

used to analyze the data after testing. The ratio of the specimen height to diameter was

not suitable for modulus determination according to the standard. Camera images were

taken of each specimen after compression testing. A macro lens was used on the camera

to take close-up images of the top and sides of each specimen. The top surface images

were converted to binary where the top surface was black and the background, white. FIJI

software was used to calculate the aspect ratio of the specimens using the Feret

diameters. The Feret diameters are the longest and smallest diameters of the specimen.

The longest/smallest diameter gives the aspect ratio from the plastic deformation.

Page 61: Development of lightweight materials by meso- and

47

After testing, the samples were prepared for metallography. They were mounted

in bakelite and polished to 0.05 microns in colloidal silica. For optical microscopy, the

samples were electrolytically etched in a 60:40 nitric acid: water solution at 6 volts for

approximately 5 seconds. Vicker’s hardness measurements were done on the top surface

of a tested specimen after the images were taken. This was done in a grid to map the

hardness across the deformed specimen. This tracked the local magnitude of the plastic

deformation.

4. RESULTS AND DISCUSSION

4.1. EXPERIMENTAL WORK

The as-built sample density was not significantly different between samples. The

samples were 98% ± 1.0% dense on average. The compression testing analysis found

0.2% offset yield strength and aspect ratio. The results for each sample are shown in

Table 2. The maximum yield strength and the lowest aspect ratio for each hatch angle are

in bold font. For each hatch angle, the cylinder built either in the [100] or [010] direction

had the highest yield strength. The aspect ratio of the 67° and 105° samples were low -

with one exception equal to or below 1.10 on average. The 0° samples, except for the

[011] sample, were noticeably anisotropic. The tensile yield strength of traditionally

manufactured, annealed 304L is 210 MPa, according to the ASM specification [32],

which significantly less than the yield strength values found in this work. Additively

manufactured materials are known to have increased yield strength, due in part to residual

stresses in the material and other microstructural factors.

Page 62: Development of lightweight materials by meso- and

48

Table 2: The yield strength, aspect ratio horizonal plane of each cylinder after testing foreach specimen are shown

Hatch Angle - Rotation Axis- Rotation Angle (Plane

Normal to Compression Direction) Aspect Ratio

0.2% Offset Yield Strength (MPa)

0°-[0101-90° (100) 1.22 ± 0.008 437 ± 4.700°-[1001-90° (010) 1.53 ± 0.052 489 ± 9.940°-[1001-0° (001) 1.65 ± 0.022 439 ± 7.94

0°-[0101-45° (011) 1.07 ± 0.026 413 ± 9.2667°-[0101-90° (100) 1.18 ± 0.025 492 ± 2.0067°-[1001-90° (010) 1.10 ± 0.042 489 ± 5.0867°-[1001-0° (001) 1.04 ± 0.015 473 ± 11.667°-[0101-45° (011) 1.05 ± 0.012 491 ± 9.89105°-[0101-90° (100) 1.04 ± 0.019 544 ± 10.7105°-[1001-90° (010) 1.03 ± 0.007 516 ± 14.24105°-[1001-0° (001) 1.06 ± 0.023 498 ± 3.72

105°-[0101-45° (011) 1.10 ± 0.022 485 ± 2.08

An aspect ratio greater than 1 indicates that anisotropic plastic deformation

occurred in the samples. This can occur if the test is performed improperly, or if non­

articulating platens are used. In this case, the articulating platens were used, and the

samples were sufficiently lubricated to reduce horizontal friction. Further, the degree of

anisotropy varied, even among specimens with similar strengths. The anisotropy was

consistent across specimens as well with an average standard deviation of 0.02. The

aspect ratio of the compression surface was not the only anisotropic plastic deformation

observed. Figure 10 shows the view of a compressed sample from the top. The shearing

of the cylinders is made obvious in the figures by the black arrows. The shearing was

most prominent in the 0° hatch angle specimens, and it seemed to be perpendicular to the

Page 63: Development of lightweight materials by meso- and

longest width of the compressed specimen. All specimens experienced the shearing

behavior, though it was less apparent in the samples with low aspect ratio. The 0°-[010]-

45° sample was notably different than the other 0° hatch angle samples in that the (001)

plane did not deform into a cylinder, but instead became rounded boxes, as shown in

Figure 11. Despite the low aspect ratio, the anisotropic plastic deformation was still

prevalent in this sample.

The Vicker’s hardness results for the undeformed samples were not statistically

different from each other. The average hardness was 216 ± 6 HV. According to the ASM

specification [24] for austenitic 304L steel, the Vicker’s hardness value should be 159

HV. The elevated hardness in the structure before testing supports the evidence that the

additively manufactured properties of the 304L material are significantly different from

traditionally manufactured material. The hardness was also tracked on the compression

surface of one of the specimens. The map of the hardness is shown in Figure 12. The

hardness values indicate that the sheared area of the specimen (noted on the figure) was

significantly harder than the compression surface.

4.2. MODEL WORK

It is clear from the literature that 3D printed structures depend on the way they are

printed. The mechanical property results of this study support that. The microstructure

and defect structure depend on build orientation and scan pattern. If the theory presented

by this work is true - that the melt pool boundary network microstructural feature

impacts the mechanical properties of the material - then the model of the MPBN may

49

Page 64: Development of lightweight materials by meso- and

50

Figure 10: Anisotropy in the 0°-[100]-0° specimens after compression testing, where the black arrows indicate the direction of shear plastic deformation and the white arrows

indicate the maximum width.

help predict strong and weak combinations in the parts. There are, of course, other

microstructural features that impact the mechanical properties of the part. These features

are slowly becoming well-documented in the literature. The features include grain size,

crystallographic orientation, and porosity distribution. The MPBN is, according to the

theory presented here, another microstructural feature that is a significant feature that can

directionally alter properties. Depending on the material, the MPBN may be more or less

influential.

Page 65: Development of lightweight materials by meso- and

51

Figure 11: The 0°-[010]-45° specimens sheared as the other 0° hatch angle specimens did, but the (001) anisotropy was not elliptical.

When the percent MPBN per slice was plotted over the length of the model, there

were two types of lines produced. Figure 13 shows an example of each one. Type One

was a pseudo-sinusoidal increase and decrease. Type Two was a relatively flat line. The

flat line indicates that the MPBN concentration was relatively constant in that orientation.

So, theoretically, the impact of the MPBN in that direction on the mechanical properties

could be consistent. The sinusoidal MPBN concentration was fluctuating between layers

semi-consistantly. The impact of the MPBN network in that direction will be,

presumably, similar to a composite with a strong and weak phase. Depending on the

microstructure, the MPBN in this direction could act similarly to a composite that has

Page 66: Development of lightweight materials by meso- and

52

Figure 12 A hardness map of the 0°-[100]-90° specimen after compression testing. Theorientation of the specimen is marked.

layers of material like corrugated cardboard. Since the material only consists of 304L

steel, the ‘composite’ is made of two ductile steels of (presumably) very similar

properties, so the composite-like structure’s impact on the properties may be minimal. In

all three hatch angles, the orientation that was guaranteed to produce a Type One line was

the orientation parallel with the build direction.

One way to analyze the data is to use a rule of mixtures approach to the strength.

That is, the higher the concentration of the ‘weak phase’ - the melt pool boundaries - the

weaker the material will be. So, since the material was compressed uniaxially, the plane

normal to the compression direction is the plane of interest. In Figure 14, the average

Page 67: Development of lightweight materials by meso- and

53

50.0%

45.0%

40.0%

0 5 .0 %PQ§ 3 0 .0%

25.0%

2 0 .0%

15.0%20% 30% 40% 50% 60% 70% 80%

Percent Position in ModelFigure 13: There were two types of lines produced when the MPBN concentration was plotted over the length of the model - flat and pseudo-sinusoidal. Examples of each are

shown.

yield strength for each sample was plotted against the concentration of the melt pool

boundaries through the plane normal to the compression direction. The negative linear

correlation is weak (R2-value of only 0.39), indicating other factors at play. However, the

negative correlation does lend some veracity to the theory presented here.

After compression, the MPBN distorted at the edges of the part. This is easiest to

see in the 0° hatch angle samples as the melt pool boundary network is stacked in parallel

lines due to the non-rotating scan pattern. The cylinders barreled out non-uniformly as

previously described. The previously smooth edges rippled as well, as shown in Figure

15. The build direction is indicated in each image of Figure 15 for context. In Figure 15.B

and C, the rippling roughly corresponds to the melt pool tracks in a single layer (B) and

Page 68: Development of lightweight materials by meso- and

54

C3

560540520500480460440420400

y = -667x + 6581 R2 = 0.39

J___ I___ I___ I___ |___ I___ I___ I___ I___ |___ I___ I___ I___ I___ |___ I___ I___ I___ I___ |___ I___ I___ I___ I___ |___ I___ I___ I___ I___ |___ I___ I___ I___ I___ |___ I___ I___ I___ L+ + + + H20.0% 22.0% 24.0% 26.0% 28.0% 30.0% 32.0% 34.0% 36.0%

[MPBN] in Plane Normal to Compression Direction

Figure 14: The yield strength vs concentration of the MPNB in the compression directionare correlated.

the layers (C), as if the features were compressed independently of the other layers. The

rippling also appeared along the melt pool track (C) possibly corresponding to individual

melt pools. This phenomenon was not limited to the edges of the cylinder, though it is

easiest to see there. The semi-independent plastic deformation of the melt pools occurred

within the bulk of the cylinder as well. This can be seen in the etched microstructure - the

MPBN is distorted similarly to the edges of the cylinder in all three figures. This

phenomenon is not limited to the 0° hatch angle samples, occurring in all the samples.

This implies that the MPBN was acting as a barrier to load transfer. Figure 16 shows a

before and after compression view of the MPBN and cellular structure. Before

compression testing, the MPBN was clearly visible in the material, and the cellular

structure appeared as imperfect hexagons or similar size. After compression, the

hexagons stretched and distorted in some places. The MPNB after compression testing

was less clearly defined.

Page 69: Development of lightweight materials by meso- and

55

A: 0°-[010]-90°

200 [jm

C: 0°-[100]-0°

Figure 15: The edges of the cylinders rippled after compression testing. The orientationof the part is marked.

Anisotropic plastic deformation of the parts, the result of which was the greater

than 1 aspect ratio of the compressed parts, was most obvious in the 0° hatch angle

Page 70: Development of lightweight materials by meso- and

56

Figure 16: A close view of the melt pool boundary and cellular structure before (A) andafter (B) compression testing.

samples. Figure 17 shows the side and top view of one compressed cylinder (0°-[100]-

0°), and Figure 12 shows the top view of a 0°-[100]-90° cylinder. Figure 18 shows the

side and top view of a 0°-[010]-90° cylinder. For the 0°-[100]-0° cylinder, the direction

of greatest width was parallel to the [100] direction and the direction of the lowest width

was the [010] direction. The 0°-[100]-0° sample had the highest aspect ratio. For the 0°-

[100]-90° specimen, the largest and smallest widths were the [100] and [001] directions,

respectively. For the 0°-[010]-90° specimen, the largest and smallest widths were the

[001] and [010] directions, respectively. These results in conjunction with the aspect ratio

data can be used to determine a ranking for these directions. The directions that caused

the most to the least plastic deformation: [100], [001], [010]. This behavior correlates to

the pseudo-sinusoidal behavior of the MPBN in those directions. Table 3 shows the

maximum and minimum peaks of the MPBN in those three directions for the 0° hatch

angle samples. The difference between the maximum and minimum MPBN in the planes

equates to the rank-ordered directions. The other hatch angles did not experience the

Page 71: Development of lightweight materials by meso- and

57

Figure 17: A 0°-[100]-0° specimen after compression testing shown from the side and the top of the cylinder. The orientation of each image is marked.

anisotropic behavior to the same degree as the 0° hatch angle samples. For the 67° hatch

angle samples, the three planes of interest had a low difference between the peaks, which

may explain why they did not experience much of an aspect ratio difference. For the 105°

hatch angle samples, very few planes exhibited the pseudo-sinusoidal behavior. The

maximum amplitude of the pseudo-sinusoidal behavior seems to be correlated with the

deformation in that direction.

Page 72: Development of lightweight materials by meso- and

58

Figure 18: A 0°-[010]-90° specimen after compression testing shown from the side and the top of the cylinder. The orientation of each image is marked.

5. CONCLUSIONS

In this work, anisotropic deformation in SLM 304L was investigated using a

model of the melt pool boundary network. The melt pool boundary network is a thin

phase of material that forms from the first material to solidify after the laser moves away

Page 73: Development of lightweight materials by meso- and

59

Table 3: The maximum, minimum, and difference between the pseudo-sinusoidal peaksfor the three directions of interest.

Hatch Angle Direction

Percent Maximum [MPBN1 in Plane

Normal to Direction

Percent Minimum

[MPBN1 in Plane Normal to Direction

Difference Between

Maximum and Minimum

0° [1001 68% 9% 59%0° [0011 44% 2 1 % 23%0° [0101 31% 18% 13%67° [1001 30% 26% 2 2%67° [0011 46% 24% 8%67° [0101 31% 24% 7%105° [1001 39% 15% 23%105° [0011 N/A N/A N/A105° [0101 N/A N/A N/A

from an area. The MPBN was modeled using a representative volume element containing

six layers of material. Three models were created with different hatch angles of 0°, 67°,

and 105°. The concentration of the MPBN in each model was tracked across multiple

planes. Experimental samples were printed using the same three hatch angles. Four build

directions were printed for each hatch angle. The samples were tested in compression.

The yield strength of these materials increased with increasing hatch angle. The build

orientations with the highest yield strengths were in the (0 0 1) plane as well, either built in

the [100] or [010] direction, depending on the hatch angle. The yield strength of the

samples was correlated to the concentration of the MPBN normal to the compression

direction of the experimental samples. A linearly negative correlation was found with an

R2 value of 0.39. The edges of the compressed cylinders were indented and outdented

after testing. Etched micrographs indicated that the rippling behavior occurred

corresponds to the melt pools. The compression tested samples were observed to have

Page 74: Development of lightweight materials by meso- and

local anisotropic plastic deformation, particularly in the 0° hatch angle samples. The

aspect ratio was used to quantify the degree of anisotropy per sample. The change in

hardness across the compression surface of one specimen confirmed the presence of

anisotropic plastic deformation. The MPBN concentration exhibited a pseudo-sinusoidal

behavior in the planes related to this anisotropic deformation. The amplitude of the

pseudo-sinusoidal curve seems to be related to the extent of the deformation experienced

by the specimens in that direction.

60

REFERENCES

[1] G. N. Levy, R. Schindel, and J. P. Kruth, “Rapid Manufacturing and Rapid Tooling With Layer Manufacturing (Lm) Technologies, State of the Art and Future Perspectives,” CIRP Ann. - Manuf. Technol., vol. 52, no. 2, pp. 589-609, 2003.

[2] R. Huang et al., “Energy and emissions saving potential of additive manufacturing: the case of lightweight aircraft components,” J. Clean. Prod., vol. 135, pp. 1559­1570, 2016.

[3] L. E. Murr et al., “Fabrication of Metal and Alloy Components by Additive Manufacturing: Examples of 3D Materials Science,” J. Mater. Res. Technol., vol.1, no. 1, pp. 42-54, Apr. 2012.

[4] W. E. Frazier, “Metal Additive Manufacturing: A Review,” J. Mater. Eng. Perform., vol. 23, no. 6 , pp. 1917-1928, Jun. 2014.

[5] M. Mazur, M. Leary, S. Sun, M. Vcelka, D. Shidid, and M. Brandt, “Deformation and failure behaviour of Ti-6Al-4V lattice structures manufactured by selective laser melting (SLM),” Int. J. Adv. Manuf. Technol., vol. 84, no. 5-8, pp. 1391­1411, 2016.

[6] P. Kohnen, C. Haase, J. Bultmann, S. Ziegler, J. H. Schleifenbaum, and W. Bleck, “Mechanical properties and deformation behavior of additively manufactured lattice structures of stainless steel,” Mater. Des., vol. 145, pp. 205-217, May 2018.

Page 75: Development of lightweight materials by meso- and

[7]

61

C. Yan, L. Hao, A. Hussein, and D. Raymont, “Evaluations of cellular lattice structures manufactured using selective laser melting,” Int. J. Mach. Tools Manuf., vol. 62, pp. 32-38, 2012.

[8] B. Brown, W. Everhart, and J. Dinardo, “Characterization of bulk to thin wall mechanical response transition in powder bed AM,” Rapid Prototyp. J., vol. 22, no. 5, pp. 801-809, Aug. 2016.

[9] E. Chlebus, B. Kuznicka, T. Kurzynowski, and B. Dybala, “Microstructure and mechanical behaviour of Ti—6Al—7Nb alloy produced by selective laser melting,” Mater. Charact., vol. 62, no. 5, pp. 488-495, May 2011.

[10] W. Shifeng, L. Shuai, W. Qingsong, C. Yan, Z. Sheng, and S. Yusheng, “Effect of molten pool boundaries on the mechanical properties of selective laser melting parts,” J. Mater. Process. Technol., vol. 214, no. 11, pp. 2660-2667, Nov. 2014.

[11] D. Wang, C. Song, Y. Yang, and Y. Bai, “Investigation of crystal growth mechanism during selective laser melting and mechanical property characterization of 316L stainless steel parts,” Mater. Des., vol. 100, pp. 291-299, 2016.

[12] M. S. Pham, B. Dovgyy, and P. A. Hooper, “Twinning induced plasticity in austenitic stainless steel 316L made by additive manufacturing,” Mater. Sci. Eng. A, vol. 704, pp. 102-111, Sep. 2017.

[13] T. Vilaro, C. Colin, and J. D. Bartout, “As-Fabricated and Heat-Treated Microstructures of the Ti-6Al-4V Alloy Processed by Selective Laser Melting,” Metall. Mater. Trans. A, vol. 42, no. 10, pp. 3190-3199, May 2011.

[14] A. S. Wu, D. W. Brown, M. Kumar, G. F. Gallegos, and W. E. King, “An Experimental Investigation into Additive Manufacturing-Induced Residual Stresses in 316L Stainless Steel,” Metall. Mater. Trans. A, vol. 45, no. 13, pp. 6260-6270, Dec. 2014.

[15] S. A. Khairallah, A. T. Anderson, A. Rubenchik, and W. E. King, “Laser powder- bed fusion additive manufacturing: Physics of complex melt flow and formation mechanisms of pores, spatter, and denudation zones,” Acta Mater., vol. 108, pp. 36-45, Apr. 2016.

[16] L. Li, C. Lough, A. Replogle, D. Bristow, R. Landers, and E. Kinzel, “Thermal Modeling of 304L Stainless Steel Selective Laser Melting,” in Solid Freedorm Fabrication, 2017, pp. 1068-1081.

Page 76: Development of lightweight materials by meso- and

62

[17] P. Guo, B. Zou, C. Huang, and H. Gao, “Study on microstructure, mechanical properties and machinability of efficiently additive manufactured AISI 316L stainless steel by high-power direct laser deposition,” J. Mater. Process. Technol., 2017.

[18] L. Thijs, F. Verhaeghe, T. Craeghs, J. Van Humbeeck, and J.-P. Kruth, “A study of the microstructural evolution during selective laser melting of Ti-6Al-4V,” Acta Mater., vol. 58, no. 9, pp. 3303-3312, May 2010.

[19] Y. Kok et al., “Anisotropy and heterogeneity of microstructure and mechanical properties in metal additive manufacturing: A critical review,” Mater. Des., vol. 139, pp. 565-586, Feb. 2018.

[20] T. DebRoy et al., “Additive manufacturing of metallic components - Process, structure and properties,” Prog. Mater. Sci., vol. 92, pp. 112-224, Mar. 2018.

[21] B. E. Carroll, T. A. Palmer, and A. M. Beese, “Anisotropic tensile behavior of Ti- 6Al-4V components fabricated with directed energy deposition additive manufacturing,” Acta Mater., vol. 87, pp. 309-320, Apr. 2015.

[22] A. H. Maamoun, M. Elbestawi, G. K. Dosbaeva, and S. C. Veldhuis, “Thermal post-processing of AlSi10Mg parts produced by Selective Laser Melting using recycled powder,” Addit. Manuf., vol. 21, pp. 234-247, May 2018.

[23] E. Liverani, S. Toschi, L. Ceschini, and A. Fortunato, “Effect of selective laser melting (SLM) process parameters on microstructure and mechanical properties of 316L austenitic stainless steel,” J. Mater. Process. Technol., vol. 249, pp. 255-263, Nov. 2017.

[24] A. Yadollahi, M. Mahmoudi, A. Elwany, H. Doude, L. Bian, and J. C. Newman, “Effects of crack orientation and heat treatment on fatigue-crack-growth behavior of AM 17-4 PH stainless steel,” Eng. Fract. Mech., vol. 226, 2020.

[25] K. Guan, Z. Wang, M. Gao, X. Li, and X. Zeng, “Effects of processing parameters on tensile properties of selective laser melted 304 stainless steel,” Mater. Des., vol. 50, pp. 581-586, 2013.

[26] B. Song et al., “Differences in microstructure and properties between selective laser melting and traditional manufacturing for fabrication of metal parts: A review,” Frontiers of Mechanical Engineering, vol. 10, no. 2. 2015.

[27] Z. Wang, T. A. Palmer, and A. M. Beese, “Effect of processing parameters on microstructure and tensile properties of austenitic stainless steel 304L made by directed energy deposition additive manufacturing,” Acta Mater., vol. 110, pp. 226-235, May 2016.

Page 77: Development of lightweight materials by meso- and

63

[28] Y. M. Wang et al., “Additively manufactured hierarchical stainless steels with high strength and ductility,” Nat. Mater., vol. 17, no. 1, pp. 63-71, Oct. 2017.

[29] T. M. Mower and M. J. Long, “Mechanical behavior of additive manufactured, powder-bed laser-fused materials,” Mater. Sci. Eng. A, vol. 651, pp. 198-213, Jan. 2016.

[30] S. Anandan et al., “Failure In metal honeycombs manufactured by selective laser melting of 304 L stainless steel under compression,” Virtual Phys. Prototyp., vol. 14, no. 2, pp. 114-122, Apr. 2018.

[31] O. Fashanu et al., “Effect of SLM Build Parameters on the Compressive Properties of 304L Stainless Steel,” J. Manuf. Mater. Process., vol. 3, no. 2, p. 43, Jun. 2019.

[32] ASM Aerospace Specification Metals Inc, “AISI Type 304L Stainless Steel,”ASM Material Data Sheet. [Online]. Available:http://asm.matweb.com/search/SpecificMaterial.asp?bassnum=MQ304L. [Accessed: 03-Jan-2019].

Page 78: Development of lightweight materials by meso- and

64

III. OPTIMIZATION AND CHARACTERIZATION OF NOVEL INJECTION MOLDING PROCESS FOR METAL MATRIX SYNTACTIC FOAMS

Myranda Spratt; Joseph W. Newkirk

Materials Science and Engineering Department, Missouri University of Science andTechnology, Rolla, MO 65409

ABSTRACT

Metal matrix syntactic foams are particulate composites comprised of hollow or

porous particles embedded in a metal matrix. These composites are difficult to

manufacture due primarily to the lightweight, relatively fragile filler material. A method

of low-pressure injection molding metal matrix syntactic foams was developed in this

work. First, an aqueous binder was optimized for low-pressure injection molding. A

three-component simplex design of experiment was used to model the density response to

changing the ratio of water to agar to glycerin. The model predicted the maximum

density was at a binder composition (in volume percent) of 7% agar, 4% glycerin, and

89% water. Second, this binder was used to manufacture copper matrix syntactic foams

with 0, 5, 10, and 15 volume % porous silica as the filler material. The solids loading for

these compositions decreased with increasing filler material from 55 to 44 volume %,

likely due to binder filling the pores in the porous silica particles. Finally, the sample

quality after injection molding was characterized. Only 0.11 ± 0.06 volume % carbon

remained in the samples. Silica particles were well-dispersed in the samples after

sintering, and they did not appear to be fractured. The specific strength of the copper

matrix material increased with increasing porous silica additions.

Page 79: Development of lightweight materials by meso- and

65

Keywords: Metal-matrix Composites; Metal Foams; Injection molding; Molding

compounds

1. INTRODUCTION

Lightweight structural composites are in high demand in certain industries,

especially in aerospace. These materials are generally two-phase composites of some

kind. For example, an aluminum-silicon carbide composite increases the strength of the

aluminum and slightly increases the density for an overall increase in the specific

strength and stiffness. Other types of composites combine a matrix and air - foams

sacrifice strength for drastically decreasing the density of the material. Syntactic foams

combine the best of both types of composites. Metal matrix syntactic foams (MMSF) are

3-phase particulate composites where hollow particles are suspended in a metal matrix.

[1] The pores in MMSFs are encased in the hollow spheres, which allows for both tight

control over the density and pore size of the material and restricts the porosity from

interacting with the matrix material directly. Rohatgi et. al. [2] summarized the typical

compression properties of MMSF in their review paper. MMSF exhibit other properties

as well as a result of being both foams and composites. For example, Mondal et. al. [3]

found that the wear rate of an aluminum - cenosphere MMSF was comparable to an Al-

SiC composite at low loads. Dou et. al. [4] found that MMSFs might be useful

electromagnetic shielding materials.

There are several methods of manufacturing MMSFs and several challenges in

doing so. Stir casting and pressure infiltration are the most popular methods [2]. Powder

Page 80: Development of lightweight materials by meso- and

metallurgy routes such as injection molding and press and sinter are less common. The

main challenge in manufacturing these materials is in incorporating the filler into the

matrix without segregation, sphere fracture, or negative chemical interactions between

the matrix and filler. In their review of the types of hollow spheres used in the literature,

Szlancsik et. al. [5] found that the hollow particles, or filler, are usually a glass or ceramic

material, though metal particles and expanded clay have also been utilized. In particular,

fly ash cenospheres remain a popular filler as they would otherwise be a waste-product

and can withstand higher temperatures than glass fillers [2, 6 , 7], The filler materials, as

they have very low densities, tends to float in molten metal which can cause density

gradients. The hollow particles can fracture during processing as well. [8] Material

compatibility is another challenge. Most glasses, other than pure silica, will likely melt

below the sintering or melting temperature of the metal. Diffusion of the matrix material

into the filler and vice versa can also occur, which can cause chemical reactions to occur.

[9, 10] In glasses, this can also cause the melting temperature of the filler to decrease.

Oxide ceramics might not wet strongly with the matrix. The two powders should bond

strongly so that load partitioning can occur. [11] Neville and Rabiei [12] worked around

these issues by using metal hollow particles as the filler. Majlinger et al [13] studied the

effect of using multiple types of particles in the same material. Lehmhus et. al. [14]

compared hollow glass to cenosphere filler in a steel alloy (316L) matrix.

A key part of MMSF research is the development of novel techniques to improve

the manufacturing of MMSF materials. For example, Weise et al [15] and Yang et al [16]

used similar methods to ensure an even distribution of hollow particles during infiltration

of the metal matrix. They sintered the filler before infiltration - Weise et al [15] using

66

Page 81: Development of lightweight materials by meso- and

67

glass filler and Yang et al [16] using ceramic spheres. Augmentation of a manufacturing

technique is one method of working around the manufacturing weaknesses of MMSF

materials. Other researchers, such as Orbulov [17] with pressure infiltration, optimize a

manufacturing process for MMSF materials. Another method is to use a novel

manufacturing technique such as done by Shiskin et al [18]. They sputter-coated the

hollow spheres with copper then used spark plasma sintering to densify the parts. Further,

a manufacturing technique that mitigates many of the manufacturing issues can be

selected.

Injection molding is one of those techniques that need little process optimization

to effectively make MMSF materials. Hollow particle flotation is not an issue in injection

molding as the powder is held in place by a binder. Particle fracture can still be an issue,

but gentle comminution and injecting procedures can be used. Weise et. al. [19, 20] used

injection molding to produce an iron alloy matrix with hollow glass microsphere

syntactic foams. They were able to sinter the iron matrix to the point where little matrix

porosity remained. However, the glass did appear to soften and wick into residual

porosity during sintering. This is a material compatibility problem solved by using

ceramic powders rather than glass ones. Hollow ceramic powders that are small enough

to use in injection molding and of good quality are difficult to find without requesting

specialty powders. Injection molding requires fine powders to form stable slurries that

can be injected into molds. As described by Szlancsik et. al. [5], most hollow ceramic

powders are in the 1 to 10 mm size range, not the <45 |im range that is typically used for

injection molding processes.

Page 82: Development of lightweight materials by meso- and

68

Injection molding requires careful selection of powders and a compatible binding

agent. The powder shape and size will influence both sintering and slurry rheology. Fine

powder sinter better than coarse powders due to their increased surface area. This

increase in surface area may negatively affect slurry fluidity. Bimodally distributed

powders achieve higher green densities as small particles fit between larges ones and fill

space. It is mathematically appropriate to have the fine particles in the bimodal

distribution be 1/7th of the coarse particles. For MMSFs, this might mean that the matrix

particles should be, on average, 7 times smaller than the hollow particles, or vice versa.

The choice of binder will determine the binder removal and burnout procedure. The

binder usually leaves some carbon or oxides behind, so minimizing the amount of binder

needed to create a flowable slurry is necessary.

Aqueous binders leave very little material leftover after the water backbone has

been removed. One such binder is a water-based gel is comprised of agar, glycerin, and

water. [21] The basic formula for this binder is agar and water which, when heated, forms

a polysaccharide gel. The glycerin in the formula acts as a gel strengthener. The binder

has similar rheological properties to a thermoplastic according to Labropoulos et. al. [22].

It has an added benefit of being more environmentally friendly than waxy binders as

usually more than 90% of the binder is water. The general formula for this binder was the

subject of or used in several patents [23-25]. It was also produced commercially

(Honeywell Powderflo Technologies, for example). Chen et. al. [26] produced NiTi

foams with an agar binder, though they used sucrose as the gel strengthener rather than

glycerin.

Page 83: Development of lightweight materials by meso- and

The goals of this work were to develop an aqueous binder for low-pressure

injection molding of metal matrix syntactic foams and successfully manufacture a copper

matrix syntactic foam using that binder and process. The composition of the agar-

glycerin-water binder was optimized to obtain the highest green and sintered relative

density in a bronze syntactic foam. The relative density was selected as a simple way to

check the quality of the parts. This optimized binder was then used to injection mold

copper matrix syntactic foams with a porous silica filler. Four samples were made with

an increasing volume fraction of filler. The quality of the sintered samples was analyzed.

The MMSFs were tested in flexure to determine the mechanical properties and compared

to the literature.

69

2. MATERIALS AND METHODS

2.1. POWDER CHARACTERIZATION

Powder characterization was performed on the syntactic foam materials to verify

morphology, and particle size distribution. Table 1 lists the syntactic foam materials used

in this experiment with the manufacturer-supplied chemistry and density. The particle

size distribution was characterized via particle measuring and counting from SEM

images. Each powder was dusted over the surface of a carbon dot on a flat specimen

holder. The powder was then mounted into an ASPEX scanning electron microscope

(SEM). The ASPEX has an automated feature analysis (AFA) that will identify and

measure inclusions or particles. At least 10,000 particles were measured for each powder,

except the two glasses. The borosilicate glass and porous silica powders could not be

Page 84: Development of lightweight materials by meso- and

70

distinguished from the noise in the ASPEX and so the particle ‘features’ were instead

analyzed by manual measurement. This was done using FIJI (Fiji Is Just ImageJ)

software. The AFA included the aspect ratio of each particle, which (along with visual

inspection of the SEM images) was used to determine morphology.

Table 1: Details provided by the manufacturer on each metal and ceramic powder used inthis experiment

Material Supplier Chemistry Density(g/cc)

Borosilicate Glass MO-SCICorp.

70-85 % SiO2

10-15% B2O3

5-10% Na2O 2-5% Al2O3

0.15

Porous Silica (P-S) MO-SCICorp.

SiO2 1.475

Copper Royal Metal Powders Inc

99.8% Copper 0.06%

Hydrogen loss

8.94

Bronze Royal Metal Powders Inc

88.46% Copper 11.3% Tin

0.24%Phosphorous

8.73

2.2. BINDER DEVELOPMENT

In this study, an aqueous binder was developed using a mixture model. The model

is discussed in the calculation section of this paper. The binder components were

deionized water, agar, and glycerin. The mixture model requires 10 samples with

different binder compositions to optimize the composition. The solids loading remained

constant at 55 vol%, and the solid composition was held constant as well. The target

Page 85: Development of lightweight materials by meso- and

71

sintered sample was a 30/70 volume ratio of hollow borosilicate spheres to bronze matrix.

Specimens were made in 50 g batches by manually mixing the binder and the solid

powders. To make the binder, boiling water was mixed with the glycerin and agar in the

appropriate combinations. The mixture was stirred until all the dry ingredients dissolved

into the solution, about 5 minutes. The solid components were heated to 100°C then

mixed into the binder. Each specimen was poured into a 50*20*15 cm silicone mold.

Three repetitions were done for each sample, and all specimens were made and measured

in a completely random order. Sample post-process characterization is discussed in

Section 2.4.

A binder-only trial in the injection mold machine was done to determine the rate

of water evaporation during use. In this test, the binder components were added to the

reservoir. The reservoir was then heated, while mixing, to 80°C. Five binder-only

specimens were injection molded every 30 minutes for 120 minutes. The specimens were

dried at 120°C for 12 hours. The mass of the specimens before and after drying was

measured and used to find the liquid loss of the binder. These values were then compared

over time to find the change in mass loss over time.

2.3. INJECTION MOLDING

A Peltsman MIGL-28 was used to injection mold the samples in this experiment.

This low-pressure injection mold machine includes a feedstock reservoir that allows for

compounding of the feedstock before injection molding. The reservoir is paddle mixed at

a fixed rate of 58 rpm. The reservoir temperature was set at 80°C for this experiment. Gas

pressure is applied to the reservoir during molding which pushes material through a

Page 86: Development of lightweight materials by meso- and

72

heated tube to the mold. The tube and orifice ring were set to 85°C. The gas pressure was

set at 67 kPa. The pressure was held for 10 seconds per specimen. The specimens were

demolded after approximately 30 seconds. The mold used in this experiment was a water-

cooled iron alloy rectangular mold. The dimensions were 6.21*0.99*0.66 cm.

Four compositions were injection molded in this study. Table 2 shows the volume

fraction of each component that was added to the reservoir, out of the total volume.

Thirty specimens were injection molded for each sample. Each sample was prepared by

mixing the binder on a hot plate, then adding the prepared binder to the pre-warmed IM

reservoir. The powders were added to the reservoir pre-heated to 100°C. For some

samples, extra binder was added to decrease slurry viscosity. The slurry was mixed for 1

hour after the last addition before injection molding began.

Table 2: The composition of each slurry in the injection molding reservoir before injection molding is listed, all units in volume percent

Component 0% P-S 5% P-S 10% P-S 15% P-S

Agar 3% 3% 4% 4%Glycerin 2% 2% 2% 2%Water 40% 44% 50% 50%BinderTotal

45% 49% 56% 56%

P-S 0% 3% 4% 7%Copper 55% 49% 40% 38%

Solids Total 55% 51% 44% 44%

Page 87: Development of lightweight materials by meso- and

73

2.4. SAMPLE POST-PROCESSING AND CHARACTERIZATION

After injection molding, the ‘wet’ samples were placed in a furnace at 120°C.

Samples were dried for at least 12 hours before sintering. The ‘dry’ specimens underwent

debinding and sintering in an atmospheric furnace with an attached retort. Debinding was

done at 450°C for 1 hour in flowing air. A steel plate was placed in the furnace beside the

samples to help getter escaping carbon and oxygen. Sintering was done directly after

debinding by dwelling at 450°C for 1 hour in flowing argon, then ramping to the

sintering temperature at a rate of 120 at 120°C/hour. Samples remained under flowing

argon throughout the sintering time. Bronze matrix samples were held at 900°C for 1

hour. Copper matrix samples were held at 1000°C for 10 hours. Samples were removed

from the furnace after cooling.

Characterization included density measurements and microstructural evaluation.

The geometric density of each specimen was measured after each processing stage (wet,

dry, and sintered). The geometric density calculation is mass over volume. Volume was

estimated as the length, width, and height of each specimen, measured with calipers. The

mass was measured using a laboratory balance. Sintered density was measured by

Archimedes’ method as described in ASTM C373-18 [27]. The vacuum method was

used. Theoretical density at each stage of processing was done by rule of mixtures using

the measured amounts of each component added to the injection molding reservoir for

each sample. Specimens were prepared for microstructure observation by mounting

specimens in bakelite and polishing. The ASPEX SEM was used for standardless energy-

dispersive x-ray spectroscopy (EDS) mapping of each sample. The automated feature

analysis function was used on the polished surface of one specimen from each sample as

Page 88: Development of lightweight materials by meso- and

well. This analysis measured the chemistry (using EDS), area, and average diameter of

10,000 features. The features were sorted by chemistry. The total area of the features and

the total area observed were used to calculate the total area percent of features in the

specimen.

The samples were the correct size and shape for 3-point bend testing. ASTM

C1161-18 [28] was used for this test, with test configuration B. Fifteen specimens were

tested for each sample. A fully articulating fixture was used. A strain rate of 0.5 mm/min

was used. Pre-loading was not done for these tests. Visual Basic for Excel code was used

to remove the toe region from each data set and determine the 0 .2% offset yield stress.

The 0.2% offset yield stress was calculated by first finding the slope of the elastic region.

Then, a line was created offset by 0.2% from the stress-strain curve. The intersecting

between the new line and the stress-strain curve was taken as the 0 .2% offset yield stress.

74

3. CALCULATION

The agar-glycerin-water binder used in this experiment was developed using a

mixture model. A 3-component simplex design was used. The wet, dry, sintered relative

density responses to altering the ratios of agar to glycerin to water were modeled. Ten

sample compositions are used in this model. The overall composition of each sample was

45 vol% binder, 33 vol% bronze, and 22 vol% hollow borosilicate glass. Three

replications were made for each sample. The 30 specimens were made and measured

completely randomly. JMP software was used to design this experiment and calculate the

response surfaces.

Page 89: Development of lightweight materials by meso- and

The experiment was designed by first selecting the lower bound of each

component. The lower limit of agar was determined after tests showed that agar does not

gel with sufficient strength below 4%. The glycerin was only a gel strengthener so the

lower limit was set at 0%. Finally, after about 15% agar, the agar became increasingly

difficult to completely dissolve into the water so the lower limit on the water was set at

85%. The compositions for the test samples were determined using these lower bounds.

The standard composition array is shown in Table 3. This table shows three pseudo­

components which exist at the corners of the ternary response surface. The pseudo­

components are related to the pure components by Equation 1. The ten compositions with

the pure components used in this experiment are shown in Table 4.

75

Table 3: The general composition of each sample for the model calculation, all units involume percent

Pseudo­Components

1 2 3 4 5 6 7 8 9 10

A 100% 0% 0% 50% 50% 0% 33% 67% 17% 17%B 0% 100% 0% 50% 0% 50% 33% 17% 67% 17%C 0% 0% 100% 0% 50% 50% 33% 17% 17% 67%

Table 4: The composition of each binder for the model calculation, all units in volumepercent

PureComponents

1 2 3 4 5 6 7 8 9 10

Agar 15% 4% 4% 10% 10% 4% 8% 1 1 % 6% 6%Glycerin 0% 1 1 % 0% 6% 0% 6% 4% 2% 7% 2%Water 85% 85% 96% 85% 91% 91% 89% 87% 87% 92%

Page 90: Development of lightweight materials by meso- and

76

The cubic response model (Equation 2) uses seven of the ten measured responses

to generate a response surface. Each b-value in the model corresponds to a mathematical

combination of measured responses, as shown by Equations 3 through 9. The other three

samples are used to test the lack-of-fit. Lack-of-fit measures the accuracy of the predicted

response values. The variance for the lack of fit (Slf2) was calculated by finding the

variance between the measured responses of samples 8 through 10 and the model

prediction of compositions 8 through 10. This is shown in Equation 10. The model must

also be tested to determine its viability. This was done by an F-ratio test. The selected

confidence interval for this test was 95%. The error variance (Serror2) was then calculated

with Equation 11. The F-ratio was calculated by Equation 12. This F-ratio was then

compared to the tabulated value. If the calculated F-ratio is less than the tabulated F-ratio,

then the model is considered a reasonable representation of the response surface of the

mixture within a set confidence interval. The tabulated F-ratio, in this case, has 3

specimen degrees of freedom, and 10 sample degrees of freedom.

i= iLB* ( 1 - iLB) * x t Equation 1

where i is the actual component amount (agar, glycerin, or water), Ilb is the lower bound

of component i, and Xi is the amount of the pseudo-component.

Cubic Response Model = b1x1 + b2x2 + b3x3 + b12x1x2 + b13x1x3 +

b23x2x3 + b123x1x2x3 Equation 2

where xi, X2, and X3 are the amount of pseudo-components A, B, and C, respectively, and

bi, b2, b3, bi2, bi3, b23, and bi23 are related to the measured response values as indicated in

Equations X4 to X10.

b i = 7 i Equation 3

Page 91: Development of lightweight materials by meso- and

77

b2 = y 2 Equation 4

^3 = y3 Equation 5

= 4y4 - 2 (y1 + y 2) Equation 6

= 4y5 - 2 (y1 + y3) Equation 7

= 4y6 - 2 (y2 + y3) Equation 8

* (y4 + y5 + y6) + 3(yi + y2 + y3) Equation 9

where yi to y 7 are the average measured response of samples one through seven.

s If = - y Vrd2) Equation 10

where yoi is the observed average response at the ith checkpoint (samples 8 through 10),

andypri is the predicted response at the ith checkpoint.

S2 =u erro rS p o o le d Equation 11

where Spooled2 is calculated with Equation 13 and is the pooled variance of all the data,

and n is the number of replications per sample.

F = 4 ^ -

c2 = 1 L c2‘-’pooled n _ ^ £ i *J £

Equation 12

Equation 13

where Si2 is the variance of the ith sample.

e r r o r

4. RESULTS AND DISCUSSION

4.1. POWDER CHARACTERIZATION

The particle size distributions by count, volume, cumulative count, and

cumulative volume percent for each of the four powders are shown in Figure 1. Table 5

Page 92: Development of lightweight materials by meso- and

shows the 10th, 50th, and 90th percentiles by count (D10, D50, D90, respectively) and

average for each powder. The size of the copper and porous silica powders were similar.

The borosilica and bronze powders appear to have similar sizes from the table, however,

the borosilica powder has far more particles between 10 and 20 microns than the bronze

powder as shown in Figure 1. The bronze powder was sieved by the manufacture to

remove fines, however, some less than 10 micron particles remained (~10%) which has

skewed the results. Due to the large density difference between the heavy metal powders

and the lightweight glass powders, these two composite powders will not stay mixed, if

they can be mixed at all. This can be solved by using a binder to facilitate mixing.

However, segregation while the binder is liquid will be a concern. In the future, it would

be appropriate to sieve out specific particle sizes such that the composite powders are

approximately 1:7 in size ratio. This will aid in mixing and may help prevent some

amount of segregation.

The particle size and morphology of the metal powders will affect how easy they

are to process and sinter. Irregular particles sinter marginally better than spherical

particles due to their higher surface area, but they are much less flowable. Figure 2 shows

example images of each powder that were used to determine the particle size distribution.

All the powders used here were spherical, as Figure 2 shows clearly. The size of the

copper powder is suitable for an injection molding and sinter operation. However, the

bronze powder is larger than would be typical for a sintering operation, especially a

pressure-less sintering operation as was used here. This is one reason that the bronze

powder used in the binder optimization procedure was not used for the injection molding

78

process.

Page 93: Development of lightweight materials by meso- and

79

25%

20%

100%

5%

0%

25%

20%

15%o10%

5%

0%

80%CD PLh CD>JB 3so

60%

40%

20%

0%o o o o o o o o o o o o o o o o

b Particle Diamete^^m) 100%

80%

60%

40%

20%

0%

cCDO—a>Ph0)>

"3£3O

O O O O O O O O O O O O O O O O

25% T C

20%

Particle Diameter (^m)

15%o10%

5%

0%

25%

20%

15%

10%

0 2 4 6Particle Diameter (^m)

5%

0%

100% Count Perce

80%_60%

: > 40%

: 3 20%

= O

ESSS3 VolumePercent

CumulativePercent

0%) — — Cumulativef

Volume100% Percent

80%

60%

40%

20%

0%

s<DoCDPLhCD.>J33£o

0 4 8 1216 2024 28 32 3640 44 48 52 56 60Particle Diameter (^m)

Figure 1: The particle size distribution by count and volume is shown for the four powders used in this study, a) Bronze, b) Borosilica, c) Copper, d) Porous Silica

Page 94: Development of lightweight materials by meso- and

80

Table 5: The 10th, 50th, 90th percentiles and average particle size in microns for eachpowder, all units in microns.

Powder D10 D50 D90 AverageBronze 5 56 93 57

Borosilicate 20 46 84 50Copper 3 6 13 7

P-S 8 16 29 17

Figure 2: SEM images of each powder are shown: a) Bronze, b) BorosiPorous Silica.

ica, c) Copper, d)

4.2. BINDER OPTIMIZATION

The binder development results include the cubic response equation and graph for

each measured response. Table 6 shows the b-values for each response which are applied

to Equation 2 (the cubic response equation) to calculate the response surfaces. Figure 3

Page 95: Development of lightweight materials by meso- and

81

shows response surfaces calculated from the cubic response equations for the wet (A),

dry (B), and sintered relative density (C), as well as the open porosity after sintering (D).

The area of highest relative density appears in the same general area of each graph. The

optimized composition was selected to be 7% agar, 4% glycerin, and 89% water. The

predicted values for each response given this composition are shown in the last column of

Table 6 . The error, pooled variance, lack-of-fit, and F-statistic for the mixture model are

shown in Table 7. With a 95% confidence interval, the tabulated F-statistic is 3.71. All

the F-statistics were below this value, so they are considered reasonable estimations of

the response. Because the solids loading and solid composition were kept constant across

all samples, this experiment should only reveal the trends of the binder. However, the

actual density depends strongly on the powder chemistry and particle size. This is

especially true of the sintered density. The b-values used here are valid only for this

experiment, but the trends should be relatively universal.

The binder development work was done using the bronze and hollow borosilicate

powders. During sintering, the borosilica glass reacted with the tin in the bronze alloy.

Tin oxide and silica inclusions were spotted in the bronze. This material compatibility

issue did not invalidate the binder development, as the trends found in this study are

considered to be material agnostic. This did prompt the matrix change to copper for the

injection molding portion of the study. Further, the melting temperature of the borosilica

powder was much lower than the sintering temperature of copper, so the filler material

was changed to the porous silica powder.

Page 96: Development of lightweight materials by meso- and

82

Volume Fraction AgarVolume Fraction Agar

o.87

._o 0.09

$■ 0.06

0.03

0.15 0.12 0.09 0.06

Volume Fraction Agar

0.40.45

0.50.55

0.6

Figure 3: The response surfaces for the a) wet density, b) dry density, c) bulk sintered density, d) open porosity after sintering where the legend indicates the relative density

fraction for a, b and c and the fraction of open porosity to total volume for d.

The binder used in this experiment was described by German and Bose [21] as an

aqueous gel useful for injection molding at low temperatures. The binder solidifies below

about 50°C and should be kept below 100°C (boiling) to prevent bubbles from forming in

the gel. This binder was chosen for this process specifically for several reasons. First, the

Peltsman MIGL-28 IM machine used in this study has a maximum operating temperature

of 120°C, so a relatively low-temperature binder had to be selected. Further, this binder is

Page 97: Development of lightweight materials by meso- and

easily cleaned with friction and warm water and is non-toxic. Finally, this binder is

approximately 90% water, which is both environmentally friendly and easy to remove

from the samples. Once the water is removed, the small amount of remaining binder is

removed by a binder burnout step. Though pre-made feedstock was not necessary for this

experiment, as the IM machine included a mixing apparatus as part of the machine,

industrial MIM machines may require pre-mixed feedstock. This binder is also capable of

being re-liquidized after comminution with the powders if the water has not been

removed. Scale-up of this binder to industrial processes is therefore possible.

83

Table 6: The cubic model b-values calculated for each response and the predicted valuebased on the optimized composition

Response b1 b2 b3 b12 b13 b23 b123 PredictedValues

WetRelativeDensity 0.674 1.106 1.127 0.295 0.447 -0.125 3.487 118%

DryRelativeDensity 0.674 1.051 1.148 0.350 0.393 0.136 4.201 123%SinteredRelativeDensity 0.587 0.777 0.838 0.368 0.367 0.127 2.197 91.6%Open

Porosity0.617 0.482 0.455 -0.276 -0.321 -0.086 -1.207 40.0%

Figure 4 shows the amount of water over time in binder-only injection molded

samples. During the first 10 to 20 minutes of the trial, the specimens varied in

composition (not shown). The agar and glycerin were added to the water cold, so the agar

agglomerated in the reservoir. It took approximately 30 minutes after reaching the

Page 98: Development of lightweight materials by meso- and

84

working temperature of 80°C for the agar to fully incorporate into the solution. In future

trials, the binder was mixed on a hotplate before adding it to the reservoir to prevent this

issue. The composition of the binder remained steady up to 2 hours after reaching the

working temperature. This indicates that the reservoir was sealed enough to prevent

significant amounts of water evaporation during usage. No water additions were therefore

necessary during use.

Table 7: The error for each response and relevant variance calculations are tabulated.R esponse n Spooled2 S 2Serror Sl f 2 F(calc)

Wet Relative

D ensity

3

0.029 0.010 0.002 0.159

D ry Relative

D ensity

3

0.034 0.011 0.000 0.035

Sintered Relative

D ensity

3

0.029 0.010 0.002 0.159

Open Porosity 3 0.0147 0.0049 0.0003 0.0653

4.3. INJECTION MOLDING

For the injection molding of syntactic foams, the binder is a critical part of

stopping segregation and promoting good mixing between the dissimilar powders that

make up the composite. The way this is determined is by tracking the composition of

Page 99: Development of lightweight materials by meso- and

85

Figure 4: The percent mass loss in each binder-only injection molded specimen over time

each specimen after molding. This can then be compared to the input composition of the

slurry. The volume and weight of each sample were measured three times: after injection

molding (wet), after removing all the water (dry), and after sintering (sintered). These

three measurements form the basis of the calculation to determine composition. The

appendix shows the calculation steps used to determine the composition for each of the

thirty specimens in the samples.

The porosity in the porous silica particles after sintering was not assumed to be

the same as before sintering. Instead, this value was found by measuring the average size

of the porous silica particles in the composite after sintering and comparing that to the

average size of the powder. The AFA data from the feature analysis performed on the

5%, 10%, and 15% P-S specimens were used to determine the average size of the P-S

particles after sintering. The average particle size of the porous silica particles after

sintering was 9.0 |im. Compared to the 17.4 |im average of the powder before sintering,

Page 100: Development of lightweight materials by meso- and

this change is significantly large. The particles had, on average, 45% porosity before

sintering. The difference in average size after sintering was 52%, so it appears that little

to no porosity remained intact in the porous silica spheres. There is significant error in

using this technique to estimate the porosity in the silica particles so this may not be

accurate. For example, the ASPEX tends to under-predict the size of glass materials. The

silica powder was measured manually, in contrast. For this calculation, the sphere

porosity was taken as 0% after sintering.

The results of finding the composition of each specimen are summarized in Table

8. Figures 5 and 6 show the data for each specimen. Figure 5 shows the percentage of

binder out of the total wet volume per specimen. Each specimen was assigned a number

from 1 to 30 after they were injection molded. The specimen number loosely correlates to

time, approximately 1 to 2 minutes per specimen. The binder concentration was relatively

stable over time. The amount of binder was generally low compared to what was added to

the reservoir (Table 2). The binder may have segregated from the powders in the

reservoir or during molding. Figure 6 shows the percentage of porous silica in each

specimen. The amount of porous silica was high on average, particularly in the case of

the 15% P-S sample. The filler concentration in the 5% and 10% P-S samples seemed to

fluctuate almost sinusoidally over time. The 15% P-S specimens did not exhibit this

behavior, remaining more consistent except at the ends of the graph. The compositional

fluctuations did not appear to be a segregation issue, as in that the composition would

have decreased over time. There may have been pockets of high concentrations of filler

that contributed to the fluctuating composition, which would indicate poor mixing of the

86

Page 101: Development of lightweight materials by meso- and

87

Table 8: The average volume percent of binder and porous silica in each sample afterinjection molding, as calculated

Sample % Binder % Porous Silica100% Cu 40% ± 0.6% N/A5% P -S 40% ± 2.3% 6% ± 1.9%

10% P -S 50% ± 2.3% 11% ± 3.2%15% P -S 53% ± 0.9% 24% ± 3.4%

slurry. Otherwise, it is possible that the slurry segregation was reduced or mitigated by

the constant agitation of the slurry in the IM reservoir.

The amount of matrix porosity in the wet specimens was assumed to be negligibly

small. However, three specimens from the 5% P-S sample were found to have significant

matrix porosity. These specimens were each found to have a large hole in their center,

likely due to an injection molding error. They were removed from the data, due to this

defect. The other specimens did not appear to have this issue.

4.4. BINDER BURNOUT AND SINTERING

The relative density of each sample after sintering is shown in Table 9. These are

average values calculated from the average filler volume percent of each sample in Table

8. The samples did not sinter to full density despite holding at over 90% of the melting

temperature of copper for 10 hours. This is due either to a binder burnout issue or a result

of water corrosion affecting the copper powder. Figure 7 shows an element map from a

0% P-S specimen. This image shows the distribution of copper and oxygen in the

specimen. The oxygen seems to surround the porosity, creating a shell of copper oxide.

The atmosphere used in this experiment was flowing air. A forming gas would be more

Page 102: Development of lightweight materials by meso- and

Cal

cula

ted

Bin

der

Am

ount

(V

olum

e Pe

rcen

t)

88

A 0% P-S

• 5% P-S

X 10% P-S

- 15% P-S

Specimen Number

Figure 5: The percentage of binder in each specimen versus specimen number, where specimen number is the order in which specimens were injection molded

• 5% P-S

X 10% P-S

- 15% P-S

Figure 6: The percentage of porous silica in each specimen versus specimen number, where specimen number is the order in which specimens were injection molded.

Page 103: Development of lightweight materials by meso- and

89

appropriate, as it appears the oxygen in the air reacted not just with the carbon in the

binder, but also with the copper powder. The formation of copper oxide allowed porosity

to remain in the sample during sintering.

Table 9: The relative density of each sample after each stage of sample post-processing shows the approximate amount of porosity in each sample

S a m p le

Wet Relative

Density

Dry Relative

Density

Sintered Relative

Density

1 0 0 % Cu 101% ± 1.1% 64% ± 0.5% 83% ± 1.3%

5 % P -S 99% ± 3.1% 70% ± 4.9% 88% ± 3.3%

1 0 % P -S 98% ± 2.3% 60% ± 3.2% 82% ± 2.6%

1 5 % P -S 111% ± 2.4% 70% ± 2.6% 91% ± 2.3%

*•

A, %' v,

^ ^ ©ACopper &

%

Figure 7: A element map of a 0% P-S specimen with the carbon, copper, and oxygen elements identified, the image was taken near the center of a specimen.

The amount of carbon present in the samples was found using the ASPEX AFA

analysis. One specimen from each sample was tested. The concentration of carbon was

Page 104: Development of lightweight materials by meso- and

90

0.11% ± 0.08% (volume percent) on average. This indicates that the carbon in the binder

was successfully removed from the samples.

The porous silica particles appeared to be well dispersed in the samples. Figure 8

shows a micrograph of a 15% porous silica specimen. This figure shows the that the

particles appear well dispersed. Some of the particles did appear to wick into the

surrounding porosity during sintering. Other particles were surrounded by porosity and

remained spherical. Bright spots in the silica may indicate areas of porosity as the edges

of the pores in the silica charge under the SEM. Alternatively, these could be copper that

migrated into the silica particles. The silica particle did not appear fragmented, most were

either spherical or wicked into nearby pores. The well-dispersed particles indicate good

local mixing during injection molding.

Figure 8: A micrograph of a 15% P-S specimen showing the dispersion of the porous silica particles and the morphology of the particles after sintering, the image was taken

near the center of a specimen.

Page 105: Development of lightweight materials by meso- and

91

4.5. MECHANICAL TESTING

The results of the 3-point bend testing are shown in Figure 9 compared to other

metal matrix syntactic foams. The matrix material and density for each MMSF were used

to normalize the data to better compare MMSF materials with different matrix materials.

The normalized density shows how much matrix porosity and filler porosity is in the

composite. The normalized strength is a measure of the matrix strengthening done by the

filler. For a lightweight structural material, the optimal MMSF, from a design standpoint,

should be in the upper left corner of the graph. Generally, above 1 in the strength axis and

below 1 in the density axis should be a design target. The MMSF materials made here

meet both of these targets, though they are on the low end of both targets.

5. CONCLUSIONS

Copper and porous silica (P-S) powders were used in a low-pressure injection

molding process to make rectangular specimens. Four compositions were made - 0, 5,

10, and 15 volume percent P-S with the remainder being copper. The powders used in

this experiment were expected to dry or wet mix poorly due to their similar size and

highly dissimilar densities. The binder optimized in this experiment worked well with the

syntactic foam materials. The agar-glycerin-water binder was optimized to achieve the

highest density. The optimized composition was 7% agar, 4% glycerin, and 89% water.

The specimens in each sample composition experienced some compositional fluctuations

over time. There was no strong time dependence, however, so this was unlikely to be a

segregation issue. The overall average filler concentration was higher than expected for

Page 106: Development of lightweight materials by meso- and

92

'J

Composite Density / Matrix Density

A This Work •[29]0[30]0[14]= [16]A [18]X [13]

Figure 9: The strength and density of several MMSF materials from this work and the literature [13], [14], [16], [18], [29], [30], where each composite’s strength was

normalized by the average strength of the matrix material and each composite density was normalized by the density of the matrix material for the sake of comparison. Note

that some literature values were estimated from figures to the best of the author’s ability.

each sample. The binder burnout procedure left 0.11% ± 0.08% carbon remaining in the

specimens, a reasonably low amount. The silica particles were well-dispersed in

specimens, indicating good mixing during injection molding. The yield strength of the

material was within the optimal range for a lightweight structural composite.

CONFLICT OF INTEREST

On behalf of all authors, the corresponding author states that there is no conflict

of interest.

Page 107: Development of lightweight materials by meso- and

93

APPENDIX

The concentration of each component in each specimen was required to be

calculated. The components were as follows: copper, silica, water, additives, matrix

porosity, and sphere porosity. The porous silica contained on average 45% porosity, and

this was kept separate from the matrix porosity. The data used to find these values was

the mass and volume measurements at each stage of post-processing. The stages were

wet, dry, and sintered. In each stage, the components were different. For example, there

was no water in the samples after the drying stage. Equations A1.1 to A1.6 show the

equations generated for each mass and volume measurement. The density of each

component was also assumed to be known. The density of the components was assumed

to be known, per Table 1, and assuming that the porous silica was 45% porosity, the

density of water was 1 g/cm3 and the density of the additives was also 1 g/cm3. This

calculation also assumes that the pores in the porous silica completely filled with binder

during the wet stage. To calculate the volume of sphere porosity after drying, Equations

A1.7 and A1.8 were used.

Mw = MCu + Msm + Mwater + Madd Equation A1.1

VW ^Cu + ^water + Vadd + ^Wmp Equation A1.2

Md = MCu + MSm + Madd Equation A1.3

Kd ^Cu + ^add + ^Dmp + ^Dsp Equation A1.4

Ms = MCu + MSm Equation A1.5

Vs = VCu + Vsm + Vsmp Equation A1.6

Pps = (PsmVsm')/ (VTsp + VSM) Equation A1.7

Page 108: Development of lightweight materials by meso- and

94

^Dsp ^Tsp Kv/ ( ^add + Kv) Equation A1.8

where M, V, and p represent mass, volume, and density, respectively. Subscripts W, D, S,

Cu, SM, water, and add represent wet, dry, sintered, copper, silica material, water, and

additives. Xmp andXsp represents matrix porosity and sphere porosity where X can be

W, D, S, or T (total).

It was assumed that the matrix porosity in subsequent stages depended on the

previous stage. Equations A1.9 and A1.10 show these relationships. Notably, all sphere

porosity was assumed to be converted to matrix porosity as the sphere porosity collapsed

during sintering.

^Dmp ^water + ^Wmp + ^extra Equation A1.9

^Smp ^ Dmp + ^Tsp ^ Kr/ ( ^add + Vw) + Vadd) Equation A1.10

where A represents the extent to which the matrix densified during the sintering

operation. In the 0% P-S sample, this value was found to be 18.0 ± 0.807 %. This value

was assumed to be constant across all specimens. Equations A1.1 through A1.10 were

used to calculate the concentration of each component in each injection molded

specimen.

REFERENCES

[1] N. Gupta and P. K. Rohatgi, “Metal Matrix Syntactic Foams: Processing, Microstructure, Properties and Applications,” in Metal Matrix Syntactic Foams: Processing, Microstructure, Properties and Applications, DEStech Publications, Inc, 2014, p. 370.

[2] P. K. Rohatgi, N. Gupta, B. F. Schultz, and D. D. Luong, “The Synthesis, Compressive Properties, and Applications of Metal Matrix Syntactic Foams,” JOM, vol. 63, no. 2, pp. 36-42, 2011.

Page 109: Development of lightweight materials by meso- and

[3] D. P. Mondal, S. Das, and N. Jha, “Dry sliding wear behaviour of aluminum syntactic foam,” Mater. Des., vol. 30, no. 7, pp. 2563-2568, Aug. 2009.

95

[4] Z. Dou, G. Wu, X. Huang, D. Sun, and L. Jiang, “Electromagnetic shielding effectiveness of aluminum alloy-fly ash composites,” Compos. Part A Appl. Sci. Manuf., vol. 38, no. 1, pp. 186-191, 2007.

[5] A. Szlancsik, B. Katona, A. Kemeny, and D. Karoly, “On the Filler Materials of Metal Matrix Syntactic Foams,” Materials (Basel)., vol. 12, no. 12, p. 2023, Jun. 2019.

[6] D. P. Mondal, S. Das, N. Ramakrishnan, and K. U. Bhasker, “Cenosphere filled aluminum syntactic foam made through stir-casting technique,” Compos. Part A Appl. Sci. Manuf., vol. 40, no. 3, pp. 279-288, 2009.

[7] U. S. Agrawal and S. P. Wanjari, “Physiochemical and engineering characteristics of cenosphere and its application as a lightweight construction material- A review,” Mater. Today Proc., vol. 4, no. 9, pp. 9797-9802, Jan. 2017.

[8] X. Xue and Y. Zhao, “Ti matrix syntactic foam fabricated by powder metallurgy: Particle breakage and elastic modulus,” JOM, vol. 63, no. 2, pp. 36-40, 2011.

[9] Y. Lin, Q. Zhang, and G. Wu, “Interfacial microstructure and compressive properties of Al-Mg syntactic foam reinforced with glass cenospheres,” J. Alloys Compd., vol. 655, pp. 301-308, 2016.

[10] I. Orbulov and K. Majlinger, “On the microstructure of ceramic hollow microspheres,” Period. Polytech. Mech. Eng., vol. 54, no. 2, p. 89, 2010.

[11] D. K. Balch and D. C. Dunand, “Load partitioning in aluminum syntactic foams containing ceramic microspheres,” Acta Mater., vol. 54, no. 6, pp. 1501-1511, 2006.

[12] B. P. Neville and A. Rabiei, “Composite metal foams processed through powder metallurgy,” Mater. Des., vol. 29, no. 2, pp. 388-396, 2008.

[13] K. Majlinger and I. N. Orbulov, “Characteristic compressive properties of hybrid metal matrix syntactic foams,” Mater. Sci. Eng. A, vol. 606, pp. 248-256, 2014.

[14] D. Lehmhus et al., “Quasi-static and Dynamic Mechanical Performance of Glass Microsphere- and Cenosphere-based 316L Syntactic Foams,” Procedia Mater. Sci., vol. 4, no. 0, pp. 383-387, 2014.

Page 110: Development of lightweight materials by meso- and

96

[15] J. Weise, V. Zanetti-Bueckmann, O. Yezerska, M. Schneider, and M. Haesche, “Processing, Properties and Coating of Micro-Porous Syntactic Foams,” Adv. Eng. Mater., vol. 9, no. 1-2, pp. 52-56, 2007.

[16] Q. Yang, J. Cheng, Y. Wei, B. Yu, Z. Miao, and P. Gao, “Innovative compound casting technology and mechanical properties of steel matrix syntactic foams,” J. Alloys Compd., p. 156572, Aug. 2020.

[17] I. N. Orbulov, “Metal matrix syntactic foams produced by pressure infiltration— The effect of infiltration parameters,” Mater. Sci. Eng. A, vol. 583, no. 0, pp. 11­19, 2013.

[18] A. Shishkin, M. Drozdova, V. Kozlov, I. Hussainova, and D. Lehmhus, “Vibration-Assisted Sputter Coating of Cenospheres: A New Approach for Realizing Cu-Based Metal Matrix Syntactic Foams,” Metals (Basel)., vol. 7, no. 1, p. 16, Jan. 2017.

[19] J. Weise, J. Baumeister, O. Yezerska, N. Salk, and G. B. D. Silva, “Syntactic Iron Foams with Integrated Microglass Bubbles Produced by Means of Metal Powder Injection Moulding,” Adv. Eng. Mater., vol. 12, no. 7, pp. 604-608, 2010.

[20] J. Weise, D. Lehmhus, J. Baumeister, R. Kun, M. Bayoumi, and M. Busse, “Production and Properties of 316 L Stainless Steel Cellular Materials and Syntactic Foams.,” Steel Res. Int., vol. 85, no. 3, pp. 486-497, Mar. 2014.

[21] R. M. German and A. Bose, Injection molding of metals and ceramics. Metal Powder Industries Federation, 1997.

[22] K. C. Labropoulos, S. Rangarajan, D. E. Niesz, and S. C. Danforth, “Dynamic Rheology of Agar Gel Based Aqueous Binders,” J. Am. Ceram. Soc., vol. 84, no.6, pp. 1217-1224, Dec. 2004.

[23] M. Sekido, H. Nakayama, M. Nukaya, and Y. Noro, “Injection-molding of metal or ceramic powders,” 14-May-1992.

[24] M. S. Zedalis, B. C. Sherman, and J. C. LaSalle, “Process for debinding and sintering metal injection molded parts made with an aqueous binder,” 27-Aug- 1998.

[25] A. J. Fanelli and R. D. Silvers, “Process for injection molding ceramic composition employing an agaroid gell-forming material to add green strength to a preform,” 15-May-1986.

Page 111: Development of lightweight materials by meso- and

97

[26] G. Chen, P. Cao, G. Wen, N. Edmonds, and Y. Li, “Using an agar-based binder to produce porous NiTi alloys by metal injection moulding,” Intermetallics, vol. 37, pp. 92-99, Jun. 2013.

[27] ASTM International, “C373-18 Standard Test Methods for Determination of Water Absorption and Associated Properties by Vacuum Method for Pressed Ceramic Tiles and Glass Tiles and Boil Method for Extruded Ceramic Tiles and Non-tile Fired Ceramic Whiteware Products.” ASM International, West Conshohocken,PA, 2018.

[28] ASTM International, “C1161-18 Standard Test Method for Flexural Strength of Advanced Ceramics at Ambient Temperature.” ASTM International, West Conshohocken, PA, 2018.

[29] N. Wang, X. Chen, Y. Li, Y. Liu, H. Zhang, and X. Wang, “Preparation and compressive performance of an A356 matrix syntactic foam,” Mater. Trans., vol. 59, no. 5, pp. 699-705, 2018.

[30] G. A. Rocha Rivero, B. F. Schultz, J. B. Ferguson, N. Gupta, and P. K. Rohatgi, “Compressive properties of Al-A206/SiC and Mg-AZ91/SiC syntactic foams,” J. Mater. Res., vol. 28, no. 17, pp. 2426-2435, 2013.

Page 112: Development of lightweight materials by meso- and

98

IV. INFLUENCE OF FILLER MATERIAL AND VOLUME FRACTION ON COPPER MATRIX SYNTACTIC FOAMS

Myranda Spratt, Joseph W. Newkirk

Materials Science and Engineering Department, Missouri University of Science andTechnology, Rolla, MO 65409

ABSTRACT

Syntactic foams are composite materials with hollow filler materials embedded in

the matrix. Metal matrix syntactic foams have lightweight structural applications due to

their high stiffness and strength to weight ratio. Two fillers were investigated in this

work, both with a copper matrix material. The two fillers were selected to be as different

as possible to contrast their influence on the properties of the syntactic foam. The foams

were made using injection molding with an aqueous binder. Porous silica syntactic foams

were made with 5, 10, and 15 volume percent filler, and Hollow alumina syntactic foams

were made with 10, 30, and 50 volume percent filler. Characterization of the samples

after processing found that both filler materials were compatible with the process and the

matrix material, though the porous silica powder out-performed the hollow alumina.

Mechanical testing included three-point bending and Charpy impact testing. The porous

silica improved the specific elastic properties of the copper more than the hollow alumina

particles. The alumina bubble at 10% filler maximized the impact energy absorption of

the samples tested.

Page 113: Development of lightweight materials by meso- and

99

1. INTRODUCTION

Syntactic foam materials combine two material design techniques - composites

and foams. Syntactic foams have no matrix material requirements, but the filler must be a

porosity-containing particle. The filler is usually a hollow particle, but it can also be

porous. The porous filler - usually a ceramic or glass - provides the joint benefits of

porosity and the filler material such as decreased density and increased stiffness. [1], [2]

If the filler is sufficiently bonded to the matrix, load partitioning between the matrix and

the filler will occur [3]. For that reason, MMSFs tend to perform better than metal foams

in that they have higher strength and stiffness to weight ratios [1], [4]-[8]. MMSFs are

also operational at higher temperatures than many composite materials, such as carbon

fiber composites [9]. The energy absorption capabilities of MMSF materials are quite

promising as well. The reinforcement of the porosity in the MMSF microstructure helps

to absorb energy during failure of these materials [10], [11].

Despite the properties that MMSFs show, they are not widely utilized. Syntactic

foams can be designed for specialty applications where complex behaviors are needed,

though for metal matrix syntactic foams (MMSFs) these applications are generally only

theoretical at this time. Proposed applications include lightweight armor [12], lightweight

electronic packaging (due to superior EM shielding) [13], biomedical implants (Ti-

matrix, primarily) [14], [15], and more. However, they are difficult to manufacture. The

hollow spheres used as the filler material have a necessarily low density in comparison to

the matrix material, often well below 1 g/cm3. They are therefore buoyant and float

during processing unless constrained or kept in motion. Methods of mitigating the

Page 114: Development of lightweight materials by meso- and

100

segregation issue include using binders as in injection molding [16], [17] or rapid casting

and solidification as in stir casting [11], [18], [19]. Mixing apparaus used to keep spheres

from floating can also fracture them, another potential problem. Material compatibility

with the matrix and the processing temperatures can also impede manufacturability

Often, the casting or sintering temperature of the matrix material will allow deleterious

reactions to occur to the filler. These can include the filler itself melting, corrosion of the

filler by the matrix material, or unwanted interfacial reactions [1], [4], [8], [11], [20]-[25]

Typical filler materials used in MMSF manufacturing are hollow glass spheres,

naturally occurring or by-product materials, ceramic hollow spheres, and expanded clay

particles [20], [26]-[32]. Some researchers also use metal hollow spheres [12], [33]. A

cost-benefit analysis of the more typical types of filler material was performed by

Szlancsik et al [34]. The glass spheres and naturally occurring or by-product materials

(which usually have a glassy component) are inexpensive. They are available in average

particle sizes of less than 100 microns, which can benefit may processes. However, they

tend to melt at temperatures low enough to preclude the use of high sintering or melting

temperature matrix materials such as steel [21], [35]-[37] Glasses also tend to rapidly

corrode in contact with molten metals such as aluminum [3], [31], [38]. Ceramic hollow

spheres are less likely to melt or otherwise react negatively with the matrix material.

However, the available ceramic spheres are usually above 100 microns in size, usually on

the order of millimeters [1], [34], [39]. So, the feasible choices in filler are limited to

those that will be compatible with the matrix material, the process, and will meet design

criteria.

Page 115: Development of lightweight materials by meso- and

101

Injection molding is one method of manufacturing metal matrix syntactic foams

[16], [30], [40]. This process, and powder metallurgy processes in general, are not often

utilized in the literature to make MMSF materials. Metal powders are expensive and lab-

scale injection molding machines are not widely available. There are a few major benefits

to this process, however, that merit its investigation. First, the composition of the

composite is entirely designable. Second, the binder will retard segregation of the filler

and matrix. Finally, sintering temperatures are lower than casting temperatures. High

melting temperature alloy matrix materials, or uncastable alloys, can be used with this

method. The downside is that sintering often takes much longer than the rapid cooling of

pressure infiltration or stir casting. In this method, the filler has to withstand high

temperature and densification pressures for a significant time. Many filler materials -

glass or glass-containing hollow spheres - tend to melt or diffuse into the matrix at the

required sintering temperatures. Refractory ceramic hollow particles that are available

will not typically melt at the required temperatures. However, these tend to be quite large,

which may make injection molding difficult.

In this work, two filler materials - one glass and one ceramic - were paired with a

copper matrix. The influence of the filler material on the manufacturability and properties

of the MMSF was analyzed. Low-pressure injection molding was used to manufacture

the samples. The two fillers were selected to be as different as possible - the size,

porosity distribution, and material of these fillers were on opposing extremes. The copper

matrix was selected as a high-density metal with a sintering temperature higher than

aluminum, but lower than the melting temperature of the glass filler. The high-density

Page 116: Development of lightweight materials by meso- and

metal contrasts with the low-density filler to allow small volume fractions of filler to

have a significant influence on the composite density.

102

2. EXPERIMENTAL PROCEDURE

In this experiment, copper matrix syntactic foams were prepared from powder

using two different filler materials. The gas atomized pure copper powder was

manufactured by Royal Metal Powders Inc. No sintering aids were added to this powder.

Porous silica (P-S) powder from Mo-Sci Corp was one filler material. This powder was

spherical with 45% porosity, per the manufacturer. The porosity in this powder was

distributed throughout each particle. Alumina bubble (AB) from Washington Mills was

the second filler. These particles were hollow, as is typical of syntactic foam filler

material. The density and average cell wall size of the alumina bubble particles was not

given by the manufacturer. The density was measured directly by mounting AB powder

in epoxy and grinding the spheres to approximately the center of the average particle size.

ImageJ software was used to measure the cell wall thickness and particle size of the AB

particles from SEM images. This data was then used to calculate the average density of

the AB powder. Powder characterization also included particle size distribution via

scanning electron microscope (SEM) for each material. The powders were imaged, and

then automated software was used to determine the diameter of each particle in each

image for up to 10,000 particles per powder. The powder was prepared for imaging by

brushing a small (<1g) sample of the powder over a carbon dot. The non-conductive

powders were then sputter-coated with palladium.

Page 117: Development of lightweight materials by meso- and

103

Seven syntactic foam compositions were injection molded in a Peltsman MIGL-

28 low-pressure injection molding machine. The volume percent of filler and copper of

each sample as input into the machine is listed in Table 1, as is the solids loading of each

sample. The MIGL-28 uses a reservoir with an attached paddle mixer to combine the

binder and powders. Gas pressure is applied to the reservoir which pushes the material

through a thin tube, through an orifice, and into the mold. The reservoir, tube, and orifice

are heated individually. In this experiment, the temperature was held at 85 °C for each

sample. The gas pressure was held for 30 seconds per specimen, and the specimens were

demolded after 1 to 2 minutes. A water-cooled 10 x 6.5 x 60 cm rectangular mold was

used.

Table 1: The solids loading, filler, and copper volume percent of each sample as inputinto the injection molding machine.

Sam ple % Solids Loading % Filler % Copper0% Filler 55% 0% 100%5% P-S 51% 5% 95%10% P-S 44% 10% 90%15% P-S 44% 15% 85%10% AB 54% 12% 88%30% AB 53% 24% 76%50% AB 56% 55% 45%

A water-based gel binder was used in this experiment. The composition was 7%

agar, 4% glycerin, and 89% water. The solids loading of each composition started at

60%, and was adjusted as necessary to achieve an injection moldable viscosity. The

slurry was mixed for 1 hour after the last adjustment. Thirty rectangular specimens were

Page 118: Development of lightweight materials by meso- and

injection molded from each sample. Each specimen was labeled in the order they were

injection molded.

Post-processing binder burnout and sintering steps occurred for each sample in

separate runs with the same equipment. Each specimen was measured for mass and

volume during each post-processing stage. The mass was measured using a laboratory

balance, and the volume was measured geometrically using calipers. The stages were

‘wet’, ‘dry’, and ‘sintered’. The wet stage occurred just after injection molding. In the

‘dry’ stage, the samples were placed in a 120 °C furnace for 12 hours to remove water

from the binder. The remaining binder was removed in a binder burnout step before the

sintering procedure. Binder burnout occurred at 450 °C for 1 hour under flowing air. The

samples were then ramped at 2°C/min to 1000 °C under flowing argon. They were held at

the sintering temperature for 10 hours.

The samples were characterized before mechanical testing. The extent of binder

segregation over time during injection molding was determined by calculating the mass

loss of each specimen from the wet to the sintered stage. Equation 1 was used to

determine this value for each specimen. A regression analysis was then used to determine

if the binder separated from the solids. Segregation between the filler and copper powder

was done by a regression analysis of the sintered relative density versus time. The

relative density was found by Equations 2 and 3. Microstructural analysis of the sintered

samples was done with extra specimens which were injection molded after the thirty test

specimens were made. The specimens were mounted in bakelite and ground down to the

midpoint of the width of the specimens. EDS mapping was done for the copper-porosity,

copper-silica, and copper-alumina interfaces to determine if interfacial reactions or

104

Page 119: Development of lightweight materials by meso- and

105

detectable diffusion occurred between the materials. The particle size of the porous silica

particles in the copper matrix was measured using the same automated feature analysis

used to characterize the powder.

Mioss (Mwet ^sintered)/'^wet Equation 1

where MioSS is the percent mass lost after sintering, Mwet is the mass at the wet stage, and

Msintered is the mass at the sintered stage of each specimen.

Prelative Pspecimen' Ptheoretical Equation 2

where preiative is the relative density, pspecimen is the specimen density, and ptheoreticai is the

theoretical density calculated from Equation 3.

Pstage 2 Pcomponents * V^components Equation 3

where the subscript stage indicates the wet, dry, or sintered stage of processing, the

subscript components indicates the components present in the stage of processing, and

V% indicates volume percent.

Two mechanical tests were performed on the samples. The specimens in each

sample were split into two sets of fifteen randomly. The dimensions of the samples were,

on average, 54 ± 2.7 x 8.6 ± 0.49 x 6.1 ± 0.31 mm. One set of fifteen for each sample was

tested by three-point bending. ASTM C1161 was used as a guide. The samples were

tested without machining. A fully-articulating fixture was used in this test, which was

required due to slight warpage in some of the specimens. Test configuration B was used

with a support span of 40 mm and a load span of 20 mm. A strain rate of 0.5 mm/min was

utilized for each test. The yield strength and modulus of each sample were determined by

visual basic for excel (VBA) code. The second set of fifteen specimens for each sample

was tested via notched Charpy impact testing. ASTM E23 was used as a guide. The

Page 120: Development of lightweight materials by meso- and

samples were not machined before use and did not comply with the standard size for a

Charpy specimen. They therefore cannot be truly compared to other Charpy impact tests

beyond this paper. The notch was the standard V-notch. Due to the nature of the

composite samples, the surface finish standard was not followed. The filler particle size

was larger than the surface finish requirements in both cases. Absorbed energy was

recorded for each specimen from the machine’s digital readout.

3. RESULTS AND DISCUSSION

3.1. POWDER CHARACTERIZATION

The particle size distribution of each powder is shown in Figure 1. Each bin of the

histogram counts the number of particles greater than the label and less than the next bin

size. Figure 1.a is the copper powder size distribution. The copper had an average particle

size of 7.0 ± 4.3 microns and a median particle size of 6.0 microns. Figure 1.b is the

alumina bubble histogram. The AB powder had an average particle size of 343.1 ± 225.0

microns and a median particle size of 278.9 microns. There were significant fragments of

spheres in this powder, which contributed to the data’s skew to the left. The powder had

the widest distribution of the three powders, and it had the largest average size. Figure 1.c

shows the porous silica powder size distribution. The average particle size of the P-S

powder was 12.0 ± 6.0 microns, and the median size was 11.1 microns. This powder was

approximately double the copper powder in size, and the distributions had a similar

shape. Figure 2 shows SEM images of each powder. All the powders were spherical. On

close inspection, the AB particles appear to be comprised of fine alumina particles, as

106

Page 121: Development of lightweight materials by meso- and

107

15% x

| 10% <u

fin

5%OO

0%0 1 2 3 4 5 6 7 8 9 1011 1213 1415

o o o o o o o o o o o o o o o o

15%

10%<ufin

5%o

0%

Particle Diameter (^m)c r

1 / 1 -i L M lJ l + _ L + _i_

100%80%

60%40%20%

0%

c<uo!-h<UPin<U_>

100% 80%

60%

40% _ 20% |

0% 5

C(L)O!-h(L)Pin(L)_>33ao

c(L)O!-h<u

Pin(L)_>33ao

0 2 4 6 8 1012141618202224262830

CountPercent

^ — CumulativePercent

Particle Diameter (^m)Figure 1: The particle size distribution by count percent and cumulative count percent for

each powder where a) is copper, b) is alumina bubble, and c) is porous silica.

shown in Figure 2.d. The fine alumina particles were not fully densified with significant

porosity in the cell walls. Figure 3 is one of the SEM images used to measure the average

diameter and wall thickness of the AB particles. On average, the density was 1.36 ± 0.39

g/cm3. The large standard deviation appears to arise from the fact that the wall thickness

was relatively constant (33 ± 14 |im) regardless of the particle size. Further, the density

Page 122: Development of lightweight materials by meso- and

108

100 urn100 um

1000 200um um■

Figure 2: SEM images of the three powders - a) copper, b) silica, c) and d) alumina - show the morphology and relative size of the powders.

calculation assumed that the cell walls were fully dense alumina, which was not the case.

The 1.36 g/cm3 value for density was used to determine filler volume percent for the AB

syntactic foam specimens.

3.2. SAMPLE CHARACTERIZATION

Table 2 shows the wet and sintered density and relative density of each sample.

The standard deviation for each average is included. The wet density was higher than

expected overall, which indicates that that expected amount of binder and/or filler was

Page 123: Development of lightweight materials by meso- and

109

Figure 3: The cross-sections of AB powder particles were used to determine the densityof the AB powder.

not present in the part or that the filler density was incorrect. Both options likely occurred

to some degree. The expected filler density could have changed if particles filled with

binder or copper particles during mixing or if the AB powder contained fragments. The

P-S particles, not being hollow, would retain their density upon fragmenting unless all

pores were exposed. The AB particles are hollow thin-walled porous shells made from

very fine alumina particles. The average size of the pores in the cell wall was 6 ± 7

microns, which is approximately the same size as the copper powder. So, copper particles

can also easily enter the AB particles, as could the binder. The porous silica could have

absorbed some binder as well from surface connected porosity, increasing their effective

density. The amount of binder in the samples was measurable quantity that was corrected

in the theoretical density calculation in Table 3. This was done by comparing the mass

after the wet stage and after the sintering stage. Then the actual volume percent of the

binder was found and used to calculate a new theoretical wet density for each sample.

Page 124: Development of lightweight materials by meso- and

110

This was used to find a new relative wet density. The samples are closer to 100% overall

with this correction. The remaining error is likely attributable to the filler density. In this

case, it seems the porous silica was less suceptable to filler density error than the alumina

bubble. It likely has less fragments and did not fill with copper or binder to the extent of

the AB particles.

Table 2: The density and relative density for each stage and sample are shown with theirstandard deviation.

Sam ple Wet SinteredDensity(g/cm3)

RelativeDensity

Density(g/cm3)

RelativeDensity

0%Filler

5.84 ± 0.07 109% ± 1.28%

7.44 ± 0.12 83% ± 1.29%

5% P-S 5.42 ± 0.04 111% ± 0.87%

7.49 ± 0.15 87% ± 1.77%

10% P­S

4.36 ± 0.05 104% ± 1.22%

6.59 ± 0.26 80% ± 3.14%

15% P­S

4.64 ± 0.04 115% ± 0.9%

6.44 ± 0.21 82% ± 2.71%

10%AB

5.7 ± 0.36 119% ± 7.49%

7.4 ± 0.27 92% ± 3.32%

30%AB

5.39 ± 0.13 127% ± 3.09%

6.79 ± 0.13 95% ± 1.85%

50%AB

4.32 ± 0.32 138% ± 10.08%

4.26 ± 0.2 89% ± 4.14%

Segregation during injection molding could occur over time and it could be a

relatively steady-state separation. Both the binder and the filler could separate from the

matrix during mixing. Table 3 shows the steady-state separation of the binder and solid

material for each sample, as calculated in the previous analysis. The amount of binder

that was in each sample, on average, after injection molding was less than the amount of

Page 125: Development of lightweight materials by meso- and

111

binder that was expected for all samples. A regression analysis on each sample found that

over time binder segregation occurred significantly in the samples with alumina bubble

particles. The effect size of this increase in binder amount over time was small overall.

The other samples did not show significant evidence of segregation of the binder over

time. Segregation of the filler was measured from the density of the sintered specimens

over injection molding timeThis analysis found no evidence of significant density

changes over time in all samples, indicating that the filler amount was constant. The

steady-state separation of the filler could not be measured, as there were too many

unknowns to calculate the actual volume percent of filler in each specimen.

Table 3: The volume percent of binder in each sample was used to calculate a more accurate relative density for each sample.

Input Volume % Actual Volume % New Wet Relative Binder Binder Density

0% Filler 45% 39% 100%5% P-S 49% 41% 100%10% P-S 56% 51% 96%15% P-S 56% 51% 106%10% AB 46% 32% 98%30% AB 47% 36% 110%50% AB 44% 32% 120%

The remaining filler porosity after sintering is an important metric to judge filler

quality. To measure this for the porous silica, the average size of the silica before

sintering and after was compared. The size of the particles decreased by 33%, on average.

The expected porosity in the PS particles was 45%, so 12% porosity remained in the

Page 126: Development of lightweight materials by meso- and

silica on average. Some silica particles partially melted and wicked into nearby matrix

porosity, which caused the noted decrease in porosity. The alumina bubble porosity was

determined from SEM images, one of which is shown in Figure 4. The AB particles were

filled with copper particles, likely during injection molding. On average, the AB particles

contained 35% porosity, 34% alumina, and 30% copper. As the image shows, however,

some of the AB particles were filled with copper while others contained no copper

particles. Fragments of AB particles were also present in the microstructure. The AB

particles, in general, did not collapse and retained more porosity than the porous silica

after sintering. However, the AB particles were observed to have many fragmented

particles, while the porous silica did not.

112

1000 [im

Figure 4: Some of the AB particles in the copper matrix filled with copper powder.

In Table 2, the relative density of the samples after sintering is low, between 82%

and 95%. Matrix porosity is apparent in the microstructure, confirming that the matrix

did not densify sufficiently. EDS mapping of a 0% filler specimen, shown in Figure 5,

Page 127: Development of lightweight materials by meso- and

113

shows the likely reason for the high amount of porosity. Oxygen was found surrounding

many of the pores in the specimen. This corrosion could have occurred during the water-

removal step or the binder-burnout step of the drying stage. In either case, the copper

oxide layer prevented densification of the matrix during sintering. The matrix of the

syntactic foam made in this experiment is functionally a low porosity copper foam rather

than a dense matrix. Further, Figure 5 shows the copper oxide coated some of the internal

pores in a pseudo-syntactic foam manner, creating almost in-situ ceramic hollow spheres.

The exact amount of copper oxide present in the samples is unknown, but it is not an

insignificant amount. This behavior could be further investigated in systems where the

oxide of the metal matrix is stiff and strong, such as alumina, as a method of creating

metal matrix syntactic foams without the need for added hollow spheres. For example,

Yu et. al. [41] investigated this method of making metal matrix syntactic foams with AlN

as the in-situ filler.

The filler-copper interfaces were EDS mapped in Figures 6 and 7. Figure 6 shows

the silica-copper microstructure. The EDS did not detect noticeable silica diffusion in the

copper matrix. However, the silica particles did appear to be surrounded by porosity,

possibly from the glass shrinkage that occurred. The particles in this microstructure

image did not fall out of the matrix during grinding and polishing, however, so they were

not loose. Figure 7 shows the alumina-copper interface. There was also no evidence of

alumina diffusion into the copper matrix.

Page 128: Development of lightweight materials by meso- and

114

© #

Figure 5: Elemental mapping of a 100% copper specimen shows the distribution ofcopper and oxygen in the sample.

3.3. MECHANICAL TESTING

One purpose of metal matrix composites to which they are particularly suited, at

least theoretically, is lightweight structural materials. To that purpose, the elastic

properties of the composite are of particular interest. The specific flexural modulus and

specific 0.2% offset yield strength are shown in Figure 8. The porous silica filler

substantially improved the specific yield strength of the copper matrix. The alumina

bubble filler did not improve the specific yield strength, but they also did not negatively

Page 129: Development of lightweight materials by meso- and

115

Figure 6: EDS map of the elements present in a P-S and copper specimen; elements areindicated on the figure.

affect the specific yield strength, except for the 50% AB sample. The specific modulus

was neutrally or negatively affected by the filler. The AB samples decreased in specific

modulus as the filler amount increased. The P-S samples did not significantly change in

specific modulus as the volume percent filler increased.

There are models in the literature derived for foams that can be used to investigate

the modulus property trends further. Zhu et al [42] developed a model for open-celled

foams (Equation 1).

Page 130: Development of lightweight materials by meso- and

116

Figure 7: EDS map of the elements present in an AB and copper specimen; elements areindicated on the figure.

EC = ES * ( 1009*p ) Equation 1^1 + 1.514 * p '

where Ec, Es, and p represent composite foam modulus, solid modulus, and relative

density, respectively. Because the copper matrix includes significant amounts of copper

oxide, the modulus of the copper matrix could not be taken from literature. The

experimentally determined modulus for the 100% copper sample was assumed to be

equal to Ec. Es was then calculated using the relative density of the sample. The Es of the

Page 131: Development of lightweight materials by meso- and

117

54.5

m

4o ^3.5

f t3

2.52 ± 2s

1.5s 1 1f t

0.5 4 0

i

f

T

KTO Specific Modulus

• Specific Yield

1 1

8.0

7.0

6.0bD

5.0S

4.0

3.0 o2.0

f tGO

1.0

0.0100% Cu 5% P-S 10% P-S 15% P-S 10% AB 30% AB 50% AB

Figure 8: The specific modulus and 0.2% offset yield strength for each sample, note that modulus is in GPa and yield strength is in MPa.

was found to be 53.7 GPa, which is approximately half of the expected tensile modulus

for pure copper.

The solid modulus for the composite samples was calculated using two methods.

For the first method, rule of mixtures was used to calculate the modulus from the solid

components of the composite. The alumina was assumed to have a modulus of 375 GPa,

which is the expected modulus of pure alumina. The silica was assumed to have a

modulus of 73 GPa, which is the modulus of dense SiO2 glass. In this method, the

theoretical density of the composite was used to calculate the relative density of the

material. This method of calculating the modulus assumes that the filler material assumed

some of the load, though it does not necessarily account for the internal nature of the

filler porosity. The second method used to calculate Ec was to assume that Es is equal to

the copper matrix alone. In this method, the theoretical density used to calculate the

Page 132: Development of lightweight materials by meso- and

118

% Porous Silica

Figure 9: The specific modulus of the porous silica and alumina bubble samples contrasted with a model for metal foams by Zhu et al [42].

relative density for each sample assumed the filler material had a density of zero. Figure

9 shows the results of the models when compared to the porous silica and alumina bubble

samples. The porous silica samples remained between the models. This indicates that

load partitioning to the silica did occur, though not as much as could have been. This is

likely due to the porosity surrounding the silica particles in the matrix. The alumina

Page 133: Development of lightweight materials by meso- and

119

bubble samples match reasonably with the model where Es was assumed to be equal to

the copper matrix. This indicates that little to no load transfer occurred in these samples

during the elastic part of the test.

The results of the notched Charpy impact testing can be seen in Figure 10. The

porous silica particles negatively impacted absorbed energy. The 10% AB sample

improved the absorbed energy and the specific absorbed energy significantly. The 30%

AB sample improved specific absorbed energy, and the absorbed energy was similar to

the 0% filler sample. The 50% AB sample had the lowest absorbed energy of any sample.

These results indicate that the alumina bubble particles imparted some impact resistance

to the copper matrix at low volume fraction AB. When the AB particles began to form a

continuous network (beyond 30% AB), the impact resistance decreased.

4. CONCLUSIONS

Two filler materials were used to create two sets of copper matrix syntactic

foams. The volume percent of each filler increased over three increments - 5, 10, and

15% for the porous silica filler and 10, 30, and 50% for the alumina bubble filler. A

control sample with no filler was also made. The two fillers were different in size,

material, and porosity distribution. The copper powder had an average particle size of 7.0

± 4.3 microns. The AB powder had an average particle size of 343.1 ± 225.0 microns.

Finally, the porous silica powder had an average particle size of 12.0 ± 6.0

microns. The alumina bubble cell walls were not dense, and the cell wall porosity

allowed copper particles to enter the internal porosity. This resulted in an after-sintering

Page 134: Development of lightweight materials by meso- and

120

% Porous Silica

Figure 10: The absorbed energy of each sample

inter-particle porosity of approximately 34%, down from 66%. The porous silica partially

melted and wicked into nearby porosity. This resulted in an inter-particle porosity of

approximately 12%, down from 45%. All the samples contained significant matrix

porosity with a copper oxide coating the pore surfaces. The flexural yield strength and

modulus and the Charpy impact energy were tested for all seven samples. The porous

silica increased the specific yield strength of the syntactic foam as more filler was added

to the matrix. The sample yield strength increased from 4.9 ± 0.44 MPa/g*cm-3 at 0%

filler to 7.0 ± 0.43 MPa/g*cm-3 at 15% P-S. The alumina bubble samples’ specific yield

Page 135: Development of lightweight materials by meso- and

121

did not depend on the filler amount. The specific moduli of the porous silica samples also

did not appear to significantly depend on the filler amount. The specific modulus of the

alumina bubble samples decreased with filler amount. The moduli were compared to an

open-cell foam model, using two methods of calculating the moduli. The porous silica

particles did appear to take on some of the load, according to the model. The impact

testing results showed that of the samples tested, the absorbed energy maximized at 10%

AB. The porous silica particles negatively affected absorbed energy.

ACKNOWLEDGEMENTS

This research did not receive any specific grant from funding agencies in the

public, commercial, or not-for-profit sectors.

REFERENCES

[1] P. K. Rohatgi, N. Gupta, B. F. Schultz, and D. D. Luong, “The Synthesis, Compressive Properties, and Applications of Metal Matrix Syntactic Foams,”JOM, vol. 63, no. 2, pp. 36-42, 2011.

[2] M. Labella, V. C. Shunmugasamy, O. M. Strbik, and N. Gupta, “Compressive and thermal characterization of syntactic foams containing hollow silicon carbide particles with porous shell,” J. Appl. Polym. Sci., vol. 131, no. 17, p. n/a-n/a, 2014.

[3] D. K. Balch and D. C. Dunand, “Load partitioning in aluminum syntactic foams containing ceramic microspheres,” Acta Mater., vol. 54, no. 6, pp. 1501-1511,2006.

[4] G. Castro and S. R. Nutt, “Synthesis of syntactic steel foam using gravity-fed infiltration,” Mater. Sci. Eng. A, vol. 553, pp. 89-95, 2012.

Page 136: Development of lightweight materials by meso- and

122

[5] L. P. Zhang and Y. Y. Zhao, “Mechanical Response of Al Matrix Syntactic Foams Produced by Pressure Infiltration Casting,” J. Compos. Mater., vol. 41, no. 41, p. 17, 2007.

[6] P. K. Rohatgi, J. K. Kim, N. Gupta, S. Alaraj, and A. Daoud, “Compressive characteristics of A356/fly ash cenosphere composites synthesized by pressure infiltration technique,” Compos. Part A Appl. Sci. Manuf., vol. 37, no. 3, pp. 430­437, 2006.

[7] D. P. Mondal, M. D. Goel, and S. Das, “Compressive deformation and energy absorption characteristics of closed cell aluminum-fly ash particle composite foam,” Mater. Sci. Eng. A, vol. 507, no. 1-2, pp. 102-109, 2009.

[8] G. A. Rocha Rivero, B. F. Schultz, J. B. Ferguson, N. Gupta, and P. K. Rohatgi, “Compressive properties of Al-A206/SiC and Mg-AZ91/SiC syntactic foams,” J. Mater. Res., vol. 28, no. 17, pp. 2426-2435, 2013.

[9] D. P. Mondal, N. Jha, A. Badkul, S. Das, and R. Khedle, “High temperature compressive deformation behaviour of aluminum syntactic foam,” Mater. Sci. Eng. A, vol. 534, pp. 521-529, Feb. 2012.

[10] L. C. Zou, Q. Zhang, B. J. Pang, G. H. Wu, L. T. Jiang, and H. Su, “Dynamic compressive behavior of aluminum matrix syntactic foam and its multilayer structure,” Mater. Des., vol. 45, 2013.

[11] A. Daoud, M. T. A. El-Khair, M. Abdel-Aziz, and P. Rohatgi, “Fabrication, microstructure and compressive behavior of ZC63 Mg-microballoon foam composites,” Compos. Part A Appl. Sci. Manuf., vol. 67, no. 9, pp. 1842-1853,2007.

[12] M. Garcia-Avila, M. Portanova, and A. Rabiei, “Ballistic performance of composite metal foams,” Compos. Struct., vol. 125, pp. 202-211, Jul. 2015.

[13] X.-F. X. Meng, D.-H. D. Li, X. X.-Q. Shen, and W. Liu, “Preparation and magnetic properties of nano-Ni coated cenosphere composites,” Appl. Surf. Sci., vol. 256, no. 12, pp. 3753-3756, Apr. 2010.

[14] X. Xue and Y. Zhao, “Ti matrix syntactic foam fabricated by powder metallurgy: Particle breakage and elastic modulus,” JOM, vol. 63, no. 2, pp. 36-40, 2011.

[15] X. B. Xue, Y. Y. Zhao, V. Kearns, and R. L. Williams, “Mechanical and Biological Properties of Titanium Syntactic Foams,” in TMS, 2010, vol. 2, pp. 129-136.

Page 137: Development of lightweight materials by meso- and

123

[16] J. Yu, M. Rombouts, and G. Maes, “Cracking behavior and mechanical properties of austenitic stainless steel parts produced by laser metal deposition,” Mater. Des., vol. 45, pp. 228-235, Mar. 2013.

[17] L. Peroni et al., “Investigation of the mechanical behaviour of AISI 316L stainless steel syntactic foams at different strain-rates,” Compos. Part B Eng., vol. 66, no. 0, pp. 430-442, 2014.

[18] D. P. Mondal, S. Das, N. Ramakrishnan, and K. U. Bhasker, “Cenosphere filled aluminum syntactic foam made through stir-casting technique,” Compos. Part A Appl. Sci. Manuf., vol. 40, no. 3, pp. 279-288, 2009.

[19] P. K. Rohatgi, D. Weiss, and N. Gupta, “Applications of Fly Ash in Synthesizing Low-Cost MMCs for Automotive and Other Applications,” JOM, vol. 58, no. 11, pp. 71-76, 2006.

[20] Y. Lin, Q. Zhang, and G. Wu, “Interfacial microstructure and compressive properties of Al-Mg syntactic foam reinforced with glass cenospheres,” J. Alloys Compd., vol. 655, pp. 301-308, 2016.

[21] J. Weise, J. Baumeister, O. Yezerska, N. Salk, and G. B. D. Silva, “Syntactic iron foams with integrated microglass bubbles produced by means of metal powder injection moulding,” Adv. Eng. Mater., vol. 12, no. 7, pp. 604-608, Apr. 2010.

[22] D. Lehmhus et al., “Quasi-static and Dynamic Mechanical Performance of Glass Microsphere- and Cenosphere-based 316L Syntactic Foams,” Procedia Mater. Sci., vol. 4, no. 0, pp. 383-387, 2014.

[23] R. A. Palmer, K. Gao, T. M. Doan, L. Green, and G. Cavallaro, “Pressure infiltrated syntactic foams—Process development and mechanical properties,” Mater. Sci. Eng. A, vol. 464, no. 1-2, pp. 85-92, 2007.

[24] J. A. Santa Maria, B. F. Schultz, J. B. Ferguson, and P. K. Rohatgi, “Al-Al2O3 syntactic foams - Part I: Effect of matrix strength and hollow sphere size on the quasi-static properties of Al-A206/Al2O3 syntactic foams,” Mater. Sci. Eng. A, vol. 582, no. 0, pp. 415-422, 2013.

[25] I. N. Orbulov and J. Dobranszky, “Producing metal matrix syntactic foams by pressure infiltration,” Period. Polytech. Mech. Eng., vol. 52, no. 1, pp. 35-42,2008.

[26] B. G. Kutchko and A. G. Kim, “Fly ash characterization by SEM-EDS,” Fuel, vol. 85, no. 17-18, pp. 2537-2544, 2006.

Page 138: Development of lightweight materials by meso- and

124

[27] Q. Zhang and G. Wu, “Three dimensional microstructure and image-based simulation of a fly ash/al syntactic foam using X-ray micro-computed tomography,” J. Wuhan Univ. Technol. Mater. Sci. Ed., vol. 28, no. 1, pp. 99-103, 2013.

[28] A. Daoud, M. T. A. El-Khair, A. Y. Shenouda, E. Mohammed, and P. K. Rohatgi, “Microstructure, tensile properties and electrochemical behavior of Pb alloy-45 vol.% fly ash microballoon composites,” Mater. Sci. Eng. A, vol. 526, no. 1-2, pp. 225-234, 2009.

[29] D. P. Mondal et al., “Titanium-cenosphere syntactic foam made through powder metallurgy route,” Mater. Des., vol. 34, pp. 82-89, Feb. 2012.

[30] J. Weise, J. Baumeister, O. Yezerska, N. Salk, and G. B. D. Silva, “Syntactic Iron Foams with Integrated Microglass Bubbles Produced by Means of Metal Powder Injection Moulding,” Adv. Eng. Mater., vol. 12, no. 7, pp. 604-608, 2010.

[31] D. D. Luong, V. C. Shunmugasamy, N. Gupta, D. Lehmhus, J. Weise, and J. Baumeister, “Quasi-static and high strain rates compressive response of iron and Invar matrix syntactic foams,” Mater. Des., vol. 66, pp. 516-531, Feb. 2015.

[32] D. Lehmhus et al., “Quasi-static and Dynamic Mechanical Performance of Glass Microsphere- and Cenosphere-based 316L Syntactic Foams,” Procedia Mater. Sci., vol. 4, pp. 383-387, Jan. 2014.

[33] A. Rabiei and L. J. Vendra, “A comparison of composite metal foam’s properties and other comparable metal foams,” Mater. Lett., vol. 63, no. 5, pp. 533-536, Feb.2009.

[34] A. Szlancsik, B. Katona, A. Kemeny, and D. Karoly, “On the Filler Materials of Metal Matrix Syntactic Foams,” Materials (Basel)., vol. 12, no. 12, p. 2023, Jun. 2019.

[35] L. Peroni, M. Scapin, M. Avalle, J. Weise, and D. Lehmhus, “Dynamic mechanical behavior of syntactic iron foams with glass microspheres,” Mater. Sci. Eng. A, vol. 552, pp. 364-375, Aug. 2012.

[36] L. Peroni et al., “High Strain Rate Tensile and Compressive Testing and Performance of Mesoporous Invar (FeNi36) Matrix Syntactic Foams Produced by Feedstock Extrusion,” Adv. Eng. Mater., Nov. 2016.

[37] D. Lehmhus et al., “Quasi-static and Dynamic Mechanical Performance of Glass Microsphere- and Cenosphere-based 316L Syntactic Foams,” Procedia Mater. Sci., vol. 4, pp. 383-387, Jan. 2014.

Page 139: Development of lightweight materials by meso- and

125

[38] M. Spratt, J. W. Newkirk, and K. Chandrashekhara, “Aluminum Matrix Syntactic Foam Fabricated with Additive Manufacturing,” in Solid State Symposium, 2017.

[39] L. Licitra, D. D. Luong, O. M. Strbik Iii, and N. Gupta, “Dynamic properties of alumina hollow particle filled aluminum alloy A356 matrix syntactic foams,” Mater. Des., vol. 66, Part B, no. 0, pp. 504-515, 2015.

[40] J. Weise, D. Lehmhus, J. Baumeister, R. Kun, M. Bayoumi, and M. Busse, “Production and Properties of 316 L Stainless Steel Cellular Materials and Syntactic Foams.,” Steel Res. Int., vol. 85, no. 3, pp. 486-497, Mar. 2014.

[41] P. Yu and M. Qian, “Metal injection moulding of in-situ formed AlN hollow sphere reinforced Al matrix syntactic foam parts,” Mater. Chem. Phys., vol. 137, no. 2, pp. 435-438, Dec. 2012.

[42] H. X. Zhu, J. F. Knott, and N. J. Mills, “Analysis of the elastic properties of open­cell foams with tetrakaidecahedral cells,” J. Mech. Phys. Solids, vol. 45, no. 3, pp. 319-343, Mar. 1997.

Page 140: Development of lightweight materials by meso- and

126

SECTION

2. CONCLUSIONS AND FUTURE WORK

In this dissertation, lightweight structures were designed, fabricated,

characterized, and their mechanical properties evaluated. A honeycomb lattice structure

and copper matrix syntactic foams were focused on in this work. The honeycomb lattice

structure is a 2D periodic cellular structure. This structure was manufactured using

powder bed selective laser melting from 304L, an austenitic stainless steel. The copper

matrix syntactic foams were manufacturing using an injection molding process. A water-

based binder was optimized for this process. Two fillers were compared to each other, a

porous silica filler and an alumina bubble filler.

The honeycomb lattice structure was evaluated in the as-built condition. Four

honeycomb samples were printed, with the wall thickness increasing from 0.2 to 0.5 mm

in 0.1 mm increments. The processing parameters used in this experiment were optimized

already to reduce porosity in the material but not for thin features. The quality of the

samples after building was analyzed by determining the extent of dimensional mismatch.

The dimensional mismatch was caused primarily by particles welded to the surfaces of

the honeycomb. The effective wall thickness of the honeycombs was increased by these

particles. The wall thickness of the honeycomb was shown to significantly affect the

strength of the material. This work was done to support model validation work by

Anandan et al [30] and Hussein et al [85].

Page 141: Development of lightweight materials by meso- and

Part of the model validation work involved finding the solid compression

properties of SLM 304L steel. During compression testing of the material, anisotropic

plastic deformation behavior was observed. Research into this phenomenon led to a

microstructure feature that has previously received little attention from the literature. This

feature, the melt pool boundary, is unique to additively manufactured material. It forms a

continuous network throughout the as-built material. This network was modeled using 3D

software to better visualize the network and calculate the concentration of the melt pool

boundary network (MPBN) in various orientations through the material. The

concentration of the MPBN was correlated to yield strength data in various build

orientations. Overall, the yield strength decreased linearly with increasing MPBN

concentration. The anisotropic plastic deformation was also linked to the concentration of

the MPBN. Directions that deformed more strongly than other directions under the same

load had higher MPBN concentration.

The lattice structure work could be extended in several ways. Property predication

via finite element analysis of the honeycomb structure would be improved by using the

effective wall thickness or by adding a variable wall thickness to the model. The

dimensional analysis work could be the basis for this improvement. This work also

highlights an important way of improving dimensional accuracy for thin-walled

structures. For thin features, the melt pool width and overlap set the strut size or wall

thickness that can be printed. This limit can be applied to future lattice structure unit

cells. The melt pool boundary network analysis can be extended to other materials and

scan pattern to see if the correlations found here hold true.

127

Page 142: Development of lightweight materials by meso- and

128

Copper matrix syntactic foams were binder injected molded with a custom binder.

The binder was a water-based gel. The binder composition was optimized using a 3-

component mixture model for maximum relative density. The optimized composition was

7% agar, 4% glycerin, and 89% water. Three samples were made with 5, 10 and 15

volume percent porous silica filler. A 0% filler sample was also made. The copper matrix

syntactic foams did not show significant evidence of segregation during injection

molding, and the binder was removed with little remaining carbon (0.11% ± 0.08%). The

matrix porosity in the samples was high (>10%) due to oxidation of the copper at the pore

surfaces. The mechanical properties of the syntactic foams increased with decreasing

density, which is ideal for these materials.

The final study compared the porous silica filler to an alumina bubble filler. The

two fillers were selected to be extremely different so as to contrast the influence they had

on the foam properties. The alumina bubble powder had an average particle size over 35

times larger than the copper powder, and the silica powder was double the size of the

copper. The porous silica spheres had distributed porosity, while the alumina particles

were thin-walled shells with porous walls. The porous shell walls allowed copper powder

to enter some alumina particles, reducing internal porosity. The copper had a sintering

temperature lower than the melting temperature of both materials. The silica did soften

during sintering, which partially collapsed some of its porosity. The alumina did not

noticeably react with the copper during sintering. The copper-alumina bubble syntactic

foams improved the impact resistance over the copper-only sample at 10% alumina

bubble. The porous silica samples did not improve impact resistance. The specific yield

Page 143: Development of lightweight materials by meso- and

strength of the porous silica increased with decreasing density. The alumina bubble’s

specific yield strength stayed relatively constant with decreasing density.

The metal matrix syntactic foam development started in this dissertation could be

extended in several ways. The binder that was developed could be used in other metal-

filler systems, in particular. This work could also be scaled-up to industrial-sized

injection molding machines. During this experiment, pores in the matrix material were

found to have an oxide coating which impeded densification. This could be a method of

making in-situ metal matrix composites. Future work could include using this binder

system to make in-situ MMSF materials with a metal matrix that has a high strength

oxide material, such as aluminum. The impact of the filler materials on different matrix

materials could also be explored.

129

Page 144: Development of lightweight materials by meso- and

130

BIBLIOGRAPHY

[1] M. F. Ashby, “The properties of foams and lattices.,” Philos. Trans. A. Math. Phys. Eng. Sci., vol. 364, no. 1838, pp. 15-30, Jan. 2006.

[2] L. Yang, D. Cormier, H. West, O. Harrysson, and K. Knowlson, “Non-stochastic Ti-6Al-4V foam structures with negative Poisson’s ratio,” Mater. Sci. Eng. A, vol. 558, no. 0, pp. 579-585, 2012.

[3] J. Banhart, “Light-Metal Foams-History of Innovation and Technological Challenges,” Adv. Eng. Mater., vol. 15, no. 3, pp. 82-111, Mar. 2013.

[4] J. Banhart, “Manufacture, characterisation and application of cellular metals and metal foams,” Prog. Mater. Sci., vol. 46, no. 6, pp. 559-632, 2001.

[5] N. Gupta and P. K. Rohatgi, Eds., Metal Matrix Syntactic Foams: Processing, Microstructure, Properties and Applications. DEStech Publications, Inc, 2014.

[6] M. F. Ashby, Materials Selection in Mechanical Design Third Edition. 2005.

[7] M. F. Ashby, “Material Property Charts,” in Materials Selection in Mechanical Design, Elsevier, 2011, pp. 57-96.

[8] W. S. Sanders and L. J. Gibson, “Mechanics of hollow sphere foams,” Mater. Sci. Eng. A, vol. 347, no. 1-2, pp. 70-85, 2003.

[9] P. K. Rohatgi, N. Gupta, B. F. Schultz, and D. D. Luong, “The Synthesis, Compressive Properties, and Applications of Metal Matrix Syntactic Foams,”JOM, vol. 63, no. 2, pp. 36-42, 2011.

[10] M. G. Rashed, M. Ashraf, R. A. W. Mines, and P. J. Hazell, “Metallic microlattice materials: A current state of the art on manufacturing, mechanical properties and applications,” Mater. Des., 2016.

[11] P. Kohnen, C. Haase, J. Bultmann, S. Ziegler, J. H. Schleifenbaum, and W. Bleck, “Mechanical properties and deformation behavior of additively manufactured lattice structures of stainless steel,” Mater. Des., vol. 145, pp. 205-217, May 2018.

[12] C. Yan, L. Hao, A. Hussein, and D. Raymont, “Evaluations of cellular lattice structures manufactured using selective laser melting,” Int. J. Mach. Tools Manuf., vol. 62, pp. 32-38, 2012.

Page 145: Development of lightweight materials by meso- and

131

[13] J. . Tu et al., “Preparation and properties of TiB2 nanoparticle reinforced copper matrix composites by in situ processing,” Mater. Lett., vol. 52, no. 6, pp. 448-452, Feb. 2002.

[14] J. P. Kruth, X. Wang, T. Laoui, and L. Froyen, “Lasers and materials in selective laser sintering,” Assem. Autom., vol. 23, no. 4, pp. 357-371, Dec. 2003.

[15] L. Thijs, F. Verhaeghe, T. Craeghs, J. Van Humbeeck, and J.-P. Kruth, “A study of the microstructural evolution during selective laser melting of Ti-6Al-4V,” Acta Mater., vol. 58, no. 9, pp. 3303-3312, May 2010.

[16] E. Chlebus, B. Kuznicka, T. Kurzynowski, and B. Dybala, “Microstructure and mechanical behaviour of Ti—6Al—7Nb alloy produced by selective laser melting,” Mater. Charact., vol. 62, no. 5, pp. 488-495, May 2011.

[17] D. Wang, C. Song, Y. Yang, and Y. Bai, “Investigation of crystal growth mechanism during selective laser melting and mechanical property characterization of 316L stainless steel parts,” Mater. Des., vol. 100, pp. 291-299, Jun. 2016.

[18] Y. M. Wang et al., “Additively manufactured hierarchical stainless steels with high strength and ductility,” Nat. Mater., vol. 17, no. 1, pp. 63-71, Oct. 2017.

[19] W. Shifeng, L. Shuai, W. Qingsong, C. Yan, Z. Sheng, and S. Yusheng, “Effect of molten pool boundaries on the mechanical properties of selective laser melting parts,” J. Mater. Process. Technol., vol. 214, no. 11, pp. 2660-2667, Nov. 2014.

[20] M. Mazur, M. Leary, S. Sun, M. Vcelka, D. Shidid, and M. Brandt, “Deformation and failure behaviour of Ti-6Al-4V lattice structures manufactured by selective laser melting (SLM),” Int. J. Adv. Manuf. Technol., vol. 84, no. 5-8, pp. 1391­1411, 2016.

[21] G. Castro and S. R. Nutt, “Synthesis of syntactic steel foam using gravity-fed infiltration,” Mater. Sci. Eng. A, vol. 553, pp. 89-95, 2012.

[22] L. P. Zhang and Y. Y. Zhao, “Mechanical Response of Al Matrix Syntactic Foams Produced by Pressure Infiltration Casting,” J. Compos. Mater., vol. 41, no. 41, p.17, 2007.

[23] P. K. Rohatgi, J. K. Kim, N. Gupta, S. Alaraj, and A. Daoud, “Compressive characteristics of A356/fly ash cenosphere composites synthesized by pressure infiltration technique,” Compos. Part A Appl. Sci. Manuf., vol. 37, no. 3, pp. 430­437, 2006.

Page 146: Development of lightweight materials by meso- and

[24] D. P. Mondal, M. D. Goel, and S. Das, “Compressive deformation and energy absorption characteristics of closed cell aluminum-fly ash particle composite foam,” Mater. Sci. Eng. A, vol. 507, no. 1-2, pp. 102-109, 2009.

132

[25] G. A. Rocha Rivero, B. F. Schultz, J. B. Ferguson, N. Gupta, and P. K. Rohatgi, “Compressive properties of Al-A206/SiC and Mg-AZ91/SiC syntactic foams,” J. Mater. Res., vol. 28, no. 17, pp. 2426-2435, 2013.

[26] A. Daoud, M. T. A. El-Khair, A. Y. Shenouda, E. Mohammed, and P. K. Rohatgi, “Microstructure, tensile properties and electrochemical behavior of Pb alloy-45 vol.% fly ash microballoon composites,” Mater. Sci. Eng. A, vol. 526, no. 1-2, pp. 225-234, 2009.

[27] M. Garcia-Avila, M. Portanova, and A. Rabiei, “Ballistic performance of composite metal foams,” Compos. Struct., vol. 125, pp. 202-211, Jul. 2015.

[28] R. A. Palmer, K. Gao, T. M. Doan, L. Green, and G. Cavallaro, “Pressure infiltrated syntactic foams—Process development and mechanical properties,” Mater. Sci. Eng. A, vol. 464, no. 1-2, pp. 85-92, 2007.

[29] A. Szlancsik, B. Katona, A. Kemeny, and D. Karoly, “On the Filler Materials of Metal Matrix Syntactic Foams,” Materials (Basel)., vol. 12, no. 12, p. 2023, Jun. 2019.

[30] S. Anandan et al., “Failure In metal honeycombs manufactured by selective laser melting of 304 L stainless steel under compression,” Virtual Phys. Prototyp., vol. 14, no. 2, pp. 114-122, Apr. 2018.

[31] M. R. Karamooz Ravari, M. Kadkhodaei, M. Badrossamay, and R. Rezaei, “Numerical investigation on mechanical properties of cellular lattice structures fabricated by fused deposition modeling,” Int. J. Mech. Sci., vol. 88, pp. 154-161, 2014.

[32] V. Br0tan, O. Fergani, K. S0rby, and T. Welo, “Mechanical Properties of Biocompatible 316L Steel Rhombic Dodecahedron Lattice Structures,” in Solid Freedom Fabrication, 2016, pp. 2087-2094.

[33] S. Josupeit, P. Delfs, D. Menge, and H. Schmid, “Manufacturability and Mechanical Characterization of Laser Sintered Lattice Structures,” pp. 2077-2086, 2016.

[34] L. Yang, “A study about size effects of 3D periodic cellular structures,” pp. 2181­2193, 2016.

Page 147: Development of lightweight materials by meso- and

133

[35] L. Yang, O. Harrysson, D. Cormier, H. West, H. Gong, and B. Stucker, “Additive Manufacturing of Metal Cellular Structures: Design and Fabrication,” JOM, vol. 67, no. 3, pp. 608-615, Mar. 2015.

[36] I. Maskery et al., “A mechanical property evaluation of graded density Al-Si10- Mg lattice structures manufactured by selective laser melting,” Mater. Sci. Eng. A, vol. 670, pp. 264-274, Jul. 2016.

[37] L. J. Gibson, “Modelling the Mechanical Behavior of Cellular Materials,” Mater. Sci. Eng. A, pp. 1-36, 1989.

[38] M. Yamashita and M. Gotoh, “Impact behavior of honeycomb structures with various cell specifications - Numerical simulation and experiment,” Int. J. Impact Eng., vol. 32, no. 1-4, pp. 618-630, Dec. 2006.

[39] P. Colombo and H. P. Degischer, “Highly porous metals and ceramics,” Mater.Sci. Technol., vol. 26, no. 10, pp. 1145-1158, 2010.

[40] Y. Zhou, H. Peng, M. Xuan, and X. Zuo, “The Microstructure and Properties of 304L Stainless Honeycombs Fabricated by Extruding and Sintering,” 2012.

[41] E. Fitzgerald and W. Everhart, “The Effect of Location on the Structure and Mechanical Properties of Selective Laser Melted 316L Stainless Steel,” Proc. 27th Annu. Int. Solid Free. Fabr. Symp., pp. 574-583, 2016.

[42] W. Everhart, P. Korinko, M. Reigel, M. Morgan, and J. Bobbitt, “CHARACTERIZATION OF SELECTIVE LASER MELTED 304L,” in MS&T, 2016.

[43] H. Bikas, P. Stavropoulos, and G. Chryssolouris, “Additive manufacturing methods and modelling approaches: a critical review,” Int. J. Adv. Manuf. Technol., vol. 83, no. 1-4, pp. 389-405, Mar. 2016.

[44] P. Hanzl, M. Zetek, T. Baksa, and T. Kroupa, “The Influence of Processing Parameters on the Mechanical Properties of SLM Parts,” Procedia Eng., vol. 100, pp. 1405-1413,Jan. 2015.

[45] L. E. Murr et al., “Fabrication of Metal and Alloy Components by Additive Manufacturing: Examples of 3D Materials Science,” J. Mater. Res. Technol., vol. 1, no. 1, pp. 42-54, Apr. 2012.

[46] L. Thijs, M. L. Montero Sistiaga, R. Wauthle, Q. Xie, J.-P. Kruth, and J. Van Humbeeck, “Strong morphological and crystallographic texture and resulting yield strength anisotropy in selective laser melted tantalum,” Acta Mater., vol. 61, no.12, pp. 4657-4668, 2013.

Page 148: Development of lightweight materials by meso- and

134

[47] J. A. Cherry, H. M. Davies, S. Mehmood, N. P. Lavery, S. G. R. Brown, and J. Sienz, “Investigation into the effect of process parameters on microstructural and physical properties of 316L stainless steel parts by selective laser melting,” Int. J. Adv. Manuf. Technol., vol. 76, no. 5-8, pp. 869-879, Feb. 2015.

[48] K. V. Yang, Y. Shi, F. Palm, X. Wu, and P. Rometsch, “Columnar to equiaxed transition in Al-Mg(-Sc)-Zr alloys produced by selective laser melting,” Scr. Mater., vol. 145, pp. 113-117, Mar. 2018.

[49] D. Dai et al., “Influence of scan strategy and molten pool configuration on microstructures and tensile properties of selective laser melting additive manufactured aluminum based parts,” Opt. Laser Technol., vol. 99, pp. 91-100, Feb. 2018.

[50] A. T. Sutton, C. S. Kriewall, M. C. Leu, and J. W. Newkirk, “Powders for Additive Manufacturing Processes: Characterization Techniques and Effects on Part Properties,” Solid Free. Fabr. Proc., pp. 1004-1030, 2016.

[51] L. Cordova, M. Campos, and T. Tinga, “POWDER CHARACTERIZATION AND OPTIMIZATION FOR ADDITIVE MANUFACTURING”

[52] L. Cordova, M. Campos, and T. Tinga, “ASSESSMENT OF MOISTURE CONTENT AND ITS INFLUENCE ON LASER BEAM MELTING FEEDSTOCK”

[53] V. Gunenthiram et al., “Experimental analysis of spatter generation and melt-pool behavior during the powder bed laser beam melting process,” J. Mater. Process. Technol., vol. 251, pp. 376-386, Jan. 2018.

[54] C. Qiu et al., “Influence of processing conditions on strut structure and compressive properties of cellular lattice structures fabricated by selective laser melting,” Mater. Sci. Eng. A, vol. 628, pp. 188-197, Mar. 2015.

[55] L. Yang et al., “Design for Additively Manufactuerd Lighweight Structure: A Perspective,” Solid Free. Fabr. Proc., vol. 6, no. 1, pp. 1-17, 2016.

[56] B. Brown, W. Everhart, and J. Dinardo, “Characterization of bulk to thin wall mechanical response transition in powder bed AM,” Rapid Prototyp. J., vol. 22, no. 5, pp. 801-809, Aug. 2016.

[57] A. Daoud, M. T. A. El-Khair, M. Abdel-Aziz, and P. Rohatgi, “Fabrication, microstructure and compressive behavior of ZC63 Mg-microballoon foam composites,” Compos. Part A Appl. Sci. Manuf., vol. 67, no. 9, pp. 1842-1853, 2007.

Page 149: Development of lightweight materials by meso- and

135

[58] A. Rabiei and L. J. Vendra, “A comparison of composite metal foam’s properties and other comparable metal foams,” Mater. Lett., vol. 63, no. 5, pp. 533-536, Feb. 2009.

[59] M. Taherishargh, I. V. Belova, G. E. Murch, and T. Fiedler, “Low-density expanded perlite-aluminium syntactic foam,” Mater. Sci. Eng. A, vol. 604, pp. 127-134, 2014.

[60] Sudarshan and M. K. Surappa, “Dry sliding wear of fly ash particle reinforced A356 Al composites,” Wear, vol. 365, no. 3-4, pp. 349-360, 2008.

[61] Sudarshan and M. K. Surappa, “Synthesis of fly ash particle reinforced A356 Al composites and their characterization,” Mater. Sci. Eng. A, vol. 480, no. 1-2, pp. 117-124, 2008.

[62] O. Couteau and D. C. Dunand, “Creep of aluminum syntactic foams,” Mater. Sci. Eng. A, vol. 488, no. 1-2, pp. 573-579, 2008.

[63] J. B. Ferguson, J. A. Santa Maria, B. F. Schultz, and P. K. Rohatgi, “Al-Al2O3 syntactic foams—Part II: Predicting mechanical properties of metal matrix syntactic foams reinforced with ceramic spheres,” Mater. Sci. Eng. A, vol. 582, no. 0, pp. 423-432, 2013.

[64] M. Y. He, M. Kiser, B. Wu, and F. W. Zok, “Influence of thermal expansion mismatch on residual stressesand flow response of microballoon composites,” Mech. Mater., vol. 23, no. 2, pp. 133-146, 1996.

[65] D. P. Mondal et al., “Titanium-cenosphere syntactic foam made through powder metallurgy route,” Mater. Des., vol. 34, pp. 82-89, Feb. 2012.

[66] J. Weise, N. Salk, U. Jehring, J. Baumeister, D. Lehmhus, and M. A. Bayoumi, “Influence of Powder Size on Production Parameters and Properties of Syntactic Invar Foams Produced by Means of Metal Powder Injection Moulding,” Adv. Eng. Mater., vol. 15, no. 3, pp. 118-122, Mar. 2013.

[67] J. A. Santa Maria, B. F. Schultz, J. B. Ferguson, and P. K. Rohatgi, “Al-Al2O3 syntactic foams - Part I: Effect of matrix strength and hollow sphere size on the quasi-static properties of Al-A206/Al2O3 syntactic foams,” Mater. Sci. Eng. A, vol. 582, no. 0, pp. 415-422, 2013.

[68] I. N. Orbulov and J. Dobranszky, “Producing metal matrix syntactic foams by pressure infiltration,” Period. Polytech. Mech. Eng., vol. 52, no. 1, pp. 35-42,2008.

Page 150: Development of lightweight materials by meso- and

136

[69] Y. Lin, Q. Zhang, and G. Wu, “Interfacial microstructure and compressiveproperties of Al-Mg syntactic foam reinforced with glass cenospheres,” J. Alloys Compd., vol. 655, pp. 301-308, 2016.

[70] J. Weise, J. Baumeister, O. Yezerska, N. Salk, and G. B. D. Silva, “Syntactic iron foams with integrated microglass bubbles produced by means of metal powder injection moulding,” Adv. Eng. Mater., vol. 12, no. 7, pp. 604-608, Apr. 2010.

[71] D. Lehmhus et al., “Quasi-static and Dynamic Mechanical Performance of Glass Microsphere- and Cenosphere-based 316L Syntactic Foams,” Procedia Mater. Sci., vol. 4, pp. 383-387, Jan. 2014.

[72] D. D. Luong, V. C. Shunmugasamy, N. Gupta, D. Lehmhus, J. Weise, and J. Baumeister, “Quasi-static and high strain rates compressive response of iron and Invar matrix syntactic foams,” Mater. Des., vol. 66, pp. 516-531, Feb. 2015.

[73] B. G. Kutchko and A. G. Kim, “Fly ash characterization by SEM-EDS,” Fuel, vol. 85, no. 17-18, pp. 2537-2544, 2006.

[74] Q. Zhang and G. Wu, “Three dimensional microstructure and image-based simulation of a fly ash/al syntactic foam using X-ray micro-computed tomography,” J. Wuhan Univ. Technol. Mater. Sci. Ed., vol. 28, no. 1, pp. 99-103, 2013.

[75] S. Mohanty and Y. P. Chugh, “Development of fly ash-based automotive brake lining,” Tribol. Int., vol. 40, no. 7, pp. 1217-1224, 2007.

[76] X. Xue and Y. Zhao, “Ti matrix syntactic foam fabricated by powder metallurgy: Particle breakage and elastic modulus,” JOM, vol. 63, no. 2, pp. 36-40, 2011.

[77] P. K. Rohatgi, R. Q. Guo, H. Iksan, E. J. Borchelt, and R. Asthana, “Pressure infiltration technique for synthesis of aluminum-fly ash particulate composite,” Mater. Sci. Eng. A, vol. 244, no. 1-2, pp. 22-30, 1998.

[78] J. Weise, J. Baumeister, O. Yezerska, N. Salk, and G. B. D. Silva, “Syntactic Iron Foams with Integrated Microglass Bubbles Produced by Means of Metal Powder Injection Moulding,” Adv. Eng. Mater., vol. 12, no. 7, pp. 604-608, 2010.

[79] J. K. Cochran, “Ceramic hollow spheres and their applications,” Curr. Opin. Solid State Mater. Sci., vol. 3, no. 5, pp. 474-479, 1998.

[80] K. Majlinger and I. N. Orbulov, “Characteristic compressive properties of hybrid metal matrix syntactic foams,” Mater. Sci. Eng. A, vol. 606, pp. 248-256, 2014.

Page 151: Development of lightweight materials by meso- and

137

[81] “Production and Properties of 316L Stainless Steel Cellular Materials and Syntactic Foams,” 2003.

[82] P. Yu and M. Qian, “Metal injection moulding of in-situ formed AlN hollow sphere reinforced Al matrix syntactic foam parts,” Mater. Chem. Phys., vol. 137, no. 2, pp. 435-438, Dec. 2012.

[83] K. Myers, P. Cortes, B. Conner, T. Wagner, B. Hetzel, and K. M. Peters,“Structure property relationship of metal matrix syntactic foams manufactured by a binder jet printing process,” Addit. Manuf., vol. 5, pp. 54-59, 2015.

[84] M. A. Ben Trad, V. Demers, R. Cote, M. Sardarian, and L. Dufresne, “Numerical simulation and experimental investigation of mold filling and segregation in low- pressure powder injection molding of metallic feedstock,” Adv. Powder Technol., vol. 31, no. 3, pp. 1349-1358, Mar. 2020.

[85] R. Hussein et al., “Effective elastic moduli of metal honeycombs manufactured using selective laser melting,” Rapid Prototyp. J., vol. 26, no. 5, pp. 971-980, Feb. 2020.

Page 152: Development of lightweight materials by meso- and

138

VITA

Myranda Shea Spratt, nee Ferris, was born in Columbia, Missoui. She earned an

Associate of Science in Pre-Engineering in May 2012 from East Central Community

College in Union, Missouri. She then went to Missouri University of Science and

Technology to earn her Bachelor of Science in Ceramic Engineering in May 2015.

Myranda accepted the Chancellor’s Distiguished Fellowship and began the Ph.D.

degree program at Missouri S&T after receiving her bachelor’s degree. Her research was

focused on additive manufacturing and lightweight composites. In December 2020,

Myranda received her Doctor of Philosophy in Materials Science and Engineering from

Missouri University of Science and Technology.