Sei sulla pagina 1di 181

Flexural Response of Syntactic Foam Core Sandwich Structures: Effects of Graded

Face Sheets and Interpenetrating Phase Composite Foam Core

by

Allen Martin Craven

A thesis submitted to the Graduate Faculty of


Auburn University
in partial fulfillment of the
requirements for the Degree of
Master of Science

Auburn, Alabama
May 9, 2011

Keywords: sandwich structures, syntactic foam, interpenetrating phase composite, three-


point bending, digital image correlation

Approved by

Hareesh V. Tippur, Chair, McWane Professor of Mechanical Engineering


Jeffrey C. Suhling, Quina Distinguished Professor of Mechanical Engineering
Ruel A. Overfelt, Professor of Materials Engineering
Abstract

In this thesis, flexural responses of sandwich structures with different syntactic

foam core architectures are studied. The syntactic foam core is made by dispersing

hollow glass microballoons in epoxy matrix at different volume fractions ranging from

20%-40%. The face sheets of the sandwich structures are made of thin AL6061 sheets.

Three types of sandwich structures, first one with a regular syntactic foam core

(identified as SFS), second one with an aluminum-syntactic foam interpenetrating core

(identified as IPC), and third one with a syntactic foam core with graded face sheet

(identified as SFS-b) are studied. Static three-point bend tests are carried out on all three

types of sandwich structures and load-deflection responses are measured. The three

different architectures are comparatively examined in terms of peak load, deflection at

failure, nonlinearity of the flexural response, and strain energy absorbed. The global

measurements are supplemented by digital image correlation measurements in the core to

map 2D deformations and strains. The measured normal and shear strain fields and

optical microscopy are used to discern failure mechanisms of the three architectures. The

SFS sandwiches fail predominantly due to face-core debonding with a large scatter in the

peak load and deflection at failure. They also show very limited nonlinearity in their

load-deflection response. The IPC core sandwiches show a substantial improvement in

the deflection at failure with a slight reduction in peak load and significant nonlinearity in

load-deflection response. The nonlinearity in this architecture is primarily due to

ii
debonding between the phases of the interpenetrating core as well as face sheet yielding.

The SFS-b sandwiches, on the other hand, show substantial improvement in both peak

load and deflection at failure along with significant nonlinearity. This is attributed to a

gradual shear strain variation at the graded face-core interface, unlike the SFS

sandwiches, resulting in face sheet yielding before failure.

The superior material under quasi-static loading based on load, deflection, and

energy metrics was the syntactic foam core sandwiches with graded face sheets (SFS-b).

The only time it might not be considered the optimal choice might be when considering

the IPC core sandwiches for their ability to use the metallic foam network to hold the

structure together longer.

Low velocity impact tests were also conducted on select IPC and SFS-b

architectures in order to further expound on the failure mechanisms in the quasi-static

case, as well as initiate dynamic characterization research. Also, the feasibility of using

DIC method in conjunction with high-speed digital photography to study impact behavior

of sandwich structures is illustrated. Contrary to the quasi-static case, IPC foam core

sandwiches appear to outperform the SFS-b sandwiches in terms of strain energy

absorption. The presence of the interpenetrating phases seems to have a positive effect

under dynamic conditions.

iii
Acknowledgements

I would first of all like to thank my advisor, Dr. Hareesh Tippur, for his expertise

and guidance throughout my graduate and undergraduate careers. His style of managing

his graduate students allowed me the freedom to complete this work in a manner that was

satisfactory to us both. Thanks also to Dr. Jeffrey Suhling and Dr. Ruel (Tony) Overfelt

for their input and willingness to serve as my committee members. I would like to thank

the U.S. Army Research Office for funding this work through grants W911NF-08-1-0285

and ARMY-W911NF-10-1-0435. I especially want to thank all of those involved in the

SMART Scholarship. This wonderful program provided me with both generous funding

for my graduate education and a civil service job in the Air Force once my degree is

complete. I consider myself lucky to have received such a wonderful award.

My time in the Failure Characterization and Optical Techniques Lab under Dr.

Tippur was aided and made easier by my colleagues. Thanks to Dr. Dongyeon Lee,

Kailash Jajam, Chandru Periasamy, Vinod Kushvaha, Robert Bedsole, and Rahul Jhaver

for their help along the way and for helping make the day to day life in the lab a

beneficial time in my life.

Finally, thanks to my family and friends. Your unwavering support and patience

were both crucial and greatly appreciated. As Warren Buffet said, “Beware of geeks

bearing formulas.”

iv
Table of Contents

Abstract............................................................................................................................ ii

Acknowledgements......................................................................................................... iv

List of Figures................................................................................................................. viii

List of Tables.................................................................................................................. xv

1. Introduction...................................................................................................... 1

1.1 Cellular Solids and Foams................................................................................... 1

1.2 Sandwich Panels with Foam Cores..................................................................... 4

1.3 Interpenetrating Phase Composites (IPCs) ......................................................... 5

1.4 Literature review ................................................................................................. 6

1.5 Sandwich Beam Theory....................................................................................... 10

1.6 Objectives............................................................................................................ 14

1.7 Organization of the Thesis................................................................................... 15

2. Material Description and Sample Preparation.......................................................... 16

2.1 Material Description............................................................................................ 16

2.1.1 Syntactic Foam.......................................................................................... 16

2.1.2 Aluminum Foam........................................................................................ 17

2.1.3 IPC Foam................................................................................................... 18

2.2 Specimen preparation ........................................................................................ 19

2.2.1 Mold Preparation....................................................................................... 19

v
2.2.2 Syntactic Foam Core Specimen Preparation............................................. 20

2.2.3 IPC Foam Core Specimen Preparation...................................................... 22

2.2.4 Syntactic Foam Core with Graded Face Sheets Specimen Preparation..... 24

2.2.5 Pattern Preparation for Optical Measurement........................................... 25

3. Flexural Characteristics of Constituents................................................................... 27

3.1 Aluminum Foam Sandwich................................................................................ 27

3.2 Syntactic Foam Sandwich .................................................................................. 32

3.3 Summary............................................................................................................. 46

4. Flexural Characteristics of IPC Foam Core Sandwich Structures............................. 48

4.1 Experimental Setup............................................................................................. 48

4.2 Quasi-Static Tests............................................................................................... 49

4.3 Summary............................................................................................................. 69

5. Flexural Characteristics of Syntactic Foam Core Sandwich Structures with Graded


Face Sheets................................................................................................................ 72

5.1 Experimental Setup............................................................................................. 72

5.2 Quasi-Static Tests............................................................................................... 73

5.3 Summary............................................................................................................. 91

6. Full Field Deformation Measurements Using Digital Image Correlation Method..... 93

6.1 Digital Image Correlation (DIC) Principle ....................................................... 93

6.2 ARAMIS™ Image Analysis Software.............................................................. 97

6.3 Experimental Setup........................................................................................... 99

6.4 Optical Measurement Results........................................................................... 100

6.4.1 Syntactic Foam (SFS) Results................................................................. 103

6.4.2 IPC Results............................................................................................... 113

vi
6.4.3 Syntactic Foam (SFS-b) Results.............................................................. 123

6.5 Impact Tests...................................................................................................... 133

7. Conclusions............................................................................................................. 144

7.1 Comparison of Results and Conclusions .......................................................... 144

7.2 Future work ...................................................................................................... 153

References..................................................................................................................... 156

Appendix A. Global/Local Displacement Comparisons.............................................. 159

Appendix B. Select ARAMIS™ Images..................................................................... 164

vii
List of Figures

Figure 1.1: Honeycomb Foam……………………………………………………… 1

Figure 1.2: Open-Celled Foam…………………………………………………….. 2

Figure 1.3: Closed-Celled Foam…………………………………………………… 2

Figure 1.4: (a) Cancellous Bone and (b) Duocel® Aluminum Foam…………….. 3

Figure 1.5: Sandwich Panels with Metallic Foam Core…………………………… 5

Figure 1.6: Idealized Interpenetrating Phase Composite Network………………... 5

Figure 1.7: Three-Point Bending Schematic of a Sandwich Beam……………….. 10

Figure 2.1: Open-Celled Aluminum Foam Sandwich Scaffold…………………… 18

Figure 2.2: Enlarged View of Duocel® Foam ……………………………………. 18

Figure 2.3: Silicone Specimen Mold and Demolded Specimen ………………….. 21

Figure 2.4: Syntactic Foam Core Sandwich Structure (SFS)……………………… 22

Figure 2.5: IPC Foam Core Sandwich Structure………………………………….. 23

Figure 2.6: Single Cast SFS-b Mold: (1) Mold, (2) AL6061 Face Sheet, and (3)
Brazed Aluminum Foam …………………………………………….. 24

Figure 2.7: Syntactic Foam Core Sandwich Structure with Graded Face Sheets
(SFS-b)……………………………………………………………….. 25

Figure 2.8: Typical Decorated Random Speckle Pattern for Optical


Measurements………………………………………………………… 26

Figure 3.1: Photograph of Three-Point Bending Test Setup……………………… 28

Figure 3.2: Camera Viewing Area………………………………………………... 29

Figure 3.3: Quasi-static Test Experimental Arrangement………………………… 30

viii
Figure 3.4: Overlay of Load-Deflection Data for 20ppi Aluminum Foam Core
Sandwich……………………………………………………………… 30

Figure 3.5: Typical Aluminum Foam Stress-Strain Response ………………….... 32

Figure 3.6: Load-Deflection Data for SFS20……………………………………... 33

Figure 3.7: Load-Deflection Data for SFS30……………………………………... 33

Figure 3.8: Load-Deflection Data for SFS40……………………………………... 34

Figure 3.9: Typical SFS Face-Core Debond……………………………………… 35

Figure 3.10: Normalized Measured Peak Load for Sandwich Structures with SF
Foam Core with Different Volume Fractions…………………………. 37

Figure 3.11: Normalized Measured Mid-Span (Load-Point) Deflection at Failure for


Sandwich Structures with SF Foam Core ……………………………. 38

Figure 3.12: Energy Absorbed for Sandwich Structures with SF Foam Core……… 40

Figure 3.13: Specific Energy Absorbed for Sandwich Structures with SF Foam
Core…………………………………………………………………… 41

Figure 3.14: (a) Failed Syntactic Foam Core Sandwich with (b) Highlighted
Face-core Debond …………………………………………………….. 42

Figure 3.15: Compressive Response of Syntactic Foam [17]………………………. 43

Figure 3.16: Predicted vs. Experimental Mid-Span Deflection, SFS20……………. 44

Figure 3.17: Predicted vs. Experimental Mid-Span Deflection, SFS30……………. 44

Figure 3.18: Predicted vs. Experimental Mid-Span Deflection, SFS40……………. 45

Figure 3.19: Overlay of Measured Load-Displacement of Sandwich Structures


with Syntactic Foam Core and Different Volume Fraction of
Microballoons in the SF ……………………………………………… 45

Figure 3.20: Face-Core Debond…………………………………………………….. 46

Figure 4.1: IPC Foam Core Sandwich Structure………………………………….. 49

Figure 4.2: Overlay of Measured Load-Deflection Data for Three Different IPC
Core Sandwich Beams (IPC20)……………………………………….. 50

ix
Figure 4.3: Overlay of Measured Load-Deflection Data for Three Different IPC
Core Sandwich Beams (IPC30)……………………………………….. 50

Figure 4.4: Overlay of Measured Load-Deflection Data for Two Different IPC
Core Sandwich Beams (IPC40)……………………………………….. 51

Figure 4.5: Typical IPC Load-Deflection Curve: Point A – Transition Point, Point
B – Crack Formation, Point C – Loss of Load Bearing Capacity ……. 53

Figure 4.6: Undeformed, P=0 N…………………………………………………... 54

Figure 4.7: Point A: Transition Point, P=1616 N ………………………………… 54

Figure 4.8: Point B: Crack Formation, P=3741 N ……………………………….. 55

Figure 4.9: Point C: Loss of Load Bearing Capacity, P=782 N ………………….. 55

Figure 4.10: Normalized Measured Peak Load for Sandwich Structures with IPC
Foam Core with Different Volume Fraction of SF in the IPC ……….. 57

Figure 4.11: Normalized Measured Mid-Span (Load-Point) Deflection at Failure


for Sandwich Structures with IPC Foam Core ……………………….. 58

Figure 4.12: Normalized Measured Load at Transition Point for Sandwich


Structures with IPC Foam Core ………………………………………. 59

Figure 4.13: Normalized Measured Mid-Span (Load-Point) Deflection at


Transition Point for Sandwich Structures with IPC Foam Core ……… 60

Figure 4.14: Energy Absorbed for Sandwich Structures with IPC Foam Core…….. 61

Figure 4.15: Specific Energy Absorbed for Sandwich Structures with IPC Foam
Core…………………………………………………………………… 63

Figure 4.16: (a) Failed IPC Foam Core Sandwich with (b) Yielded Bottom Face... 64

Figure 4.17: Magnified Images of Crack Surfaces Near the Lower Face Sheet…... 65

Figure 4.18: Compressive Response of IPC [17]…………………………………... 66

Figure 4.19: Predicted vs. Experimental Mid-Span Deflection, IPC20……………. 66

Figure 4.20: Predicted vs. Experimental Mid-Span Deflection, IPC30……………. 67

Figure 4.21: Predicted vs. Experimental Mid-Span Deflection, IPC40……………. 67

x
Figure 4.22: Overlay of Measured Load-Displacement of Sandwich Structures with
IPC Foam Core with Different Vf of Microballoons in the SF……….. 69

Figure 4.23: Illustration of IPC (a) Before and (b) After Interphase Separation
between Syntactic Foam and Aluminum Ligaments …………………. 70

Figure 5.1: Syntactic Foam Core Sandwich Structure with Graded Face Sheets
(SFS-b)………………………………………………………………… 73

Figure 5.2: Overlay of Measured Load-Deflection Data for Three Different SF


Core Sandwich Beams with Graded Face Sheets (SFSb20)…………… 74

Figure 5.3: Overlay of Measured Load-Deflection Data for Four Different SF Core
Sandwich Beams with Graded Face Sheets (SFSb30)………………… 74

Figure 5.4: Overlay of Measured Load-Deflection Data for Four Different SF Core
Sandwich Beams with Graded Face Sheets (SFSb40)………………… 75

Figure 5.5: Typical SFS-b Load-Deflection Curve: Point A – Transition Point,


Point B – Failure ……………………………………………………… 77

Figure 5.6: Undeformed, P=0 N…………………………………………………... 78

Figure 5.7: Point A: Transition Point, P=5127 N ………………………………… 78

Figure 5.8: Point B: Failure ………………………………………………………. 79

Figure 5.9: Normalized Measured Peak Load for Sandwich Structures with
Syntactic Foam Core and Graded Face Sheets ……………………….. 80

Figure 5.10: Normalized Measured Mid-Span (Load-Point) Deflection at Failure


for Sandwich Structures with Syntactic Foam Core and Graded Face
Sheets………………………………………………………………….. 81

Figure 5.11: Normalized Measured Load at Transition Point for Sandwich


Structures with Syntactic Foam Core and Graded Face Sheets……….. 82

Figure 5.12: Normalized Measured Mid-Span (Load-Point) Deflection at Transition


Point for Sandwich Structures with Syntactic Foam Core and Graded
Face Sheets ……………………………………………………………. 83

Figure 5.13: Energy Absorbed for Sandwich Structures with Syntactic Foam Core
and Graded Face Sheets……………………………………………….. 84

Figure 5.14: Specific Energy Absorbed for Sandwich Structures with Syntactic
Foam Core and Graded Face Sheets………………………………….. 86

xi
Figure 5.15: Failed SFS-b Specimen with (a) Side, (b) Top Face, and (c) Bottom
Face Views ……………………………………………………………. 87

Figure 5.16: Magnified Images (a), (b) of Failure Surfaces and (c) From Below …. 88

Figure 5.17: Predicted vs. Experimental Mid-Span Deflection, SFSb20…………... 89

Figure 5.18: Predicted vs. Experimental Mid-Span Deflection, SFSb30…………… 89

Figure 5.19: Predicted vs. Experimental Mid-Span Deflection, SFSb40…………… 90

Figure 5.20: Overlay of Measured Load-Displacement of SF Core Sandwich


Structures with Graded Face Sheets and Different Vf of Microballoons
in the SF……………………………………………………………….. 90

Figure 6.1: (a) Undeformed and (b) Deformed Speckle Images………………….. 94

Figure 6.2: Tracked Subimage for Image Correlation……………………………. 95

Figure 6.3: Coordinate System Definition………………………………………... 98

Figure 6.4: 3x3 Point Neighborhood for 2D Strain Calculation…………………. 99

Figure 6.5: Global/Local Displacement Comparison, FLEX-16…………………. 100

Figure 6.6: Strain Calibration Using PMMA…………………………………….. 102

Figure 6.7: PMMA Strain (εx) Variation Along the Height of the Beam………… 102

Figure 6.8: PMMA Shear Strain (εxy) Variation Along the Height of the Beam … 103

Figure 6.9: ARAMIS™ Results for SFS20 (FLEX-38, u)………………………… 104

Figure 6.10: ARAMIS™ Results for SFS20 (FLEX-38, εx)……………………….. 105

Figure 6.11: ARAMIS™ Results for SFS20 (FLEX-38, εxy)………………………. 106

Figure 6.12: ARAMIS™ Results for SFS30 (FLEX-39, u)………………………… 107

Figure 6.13: ARAMIS™ Results for SFS30 (FLEX-39, εx)……………………….. 108

Figure 6.14: ARAMIS™ Results for SFS30 (FLEX-39, εxy)………………………. 109

Figure 6.15: ARAMIS™ Results for SFS40 (FLEX-24, u)………………………… 110

Figure 6.16: ARAMIS™ Results for SFS40 (FLEX-24, εx)……………………….. 111

xii
Figure 6.17: ARAMIS™ Results for SFS40 (FLEX-24, εxy)………………………. 112

Figure 6.18: ARAMIS™ Results for IPC20 (FLEX-10, u)………………………… 114

Figure 6.19: ARAMIS™ Results for IPC20 (FLEX-10, εx)……………………….. 115

Figure 6.20: ARAMIS™ Results for IPC20 (FLEX-10, εxy)………………………. 116

Figure 6.21: ARAMIS™ Results for IPC30 (FLEX-16, u)………………………… 117

Figure 6.22: ARAMIS™ Results for IPC30 (FLEX-16, εx)……………………….. 118

Figure 6.23: ARAMIS™ Results for IPC30 (FLEX-16, εxy)………………………. 119

Figure 6.24: ARAMIS™ Results for IPC40 (FLEX-14, u)………………………… 120

Figure 6.25: ARAMIS™ Results for IPC40 (FLEX-14, εx)……………………….. 121

Figure 6.26: ARAMIS™ Results for IPC40 (FLEX-14, εxy)………………………. 122

Figure 6.27: ARAMIS™ Results for SFSb20 (FLEX-57, u)………………………. 124

Figure 6.28: ARAMIS™ Results for SFSb20 (FLEX-57, εx)………………………. 125

Figure 6.29: ARAMIS™ Results for SFSb20 (FLEX-57, εxy)……………………… 126

Figure 6.30: ARAMIS™ Results for SFSb30 (FLEX-59, u)………………………. 127

Figure 6.31: ARAMIS™ Results for SFSb30 (FLEX-59, εx)………………………. 128

Figure 6.32: ARAMIS™ Results for SFSb30 (FLEX-59, εxy)……………………… 129

Figure 6.33: ARAMIS™ Results for SFSb40 (FLEX-65, u)………………………. 130

Figure 6.34: ARAMIS™ Results for SFSb40 (FLEX-65, εx)………………………. 131

Figure 6.35: ARAMIS™ Results for SFSb40 (FLEX-65, εxy)……………………… 132

Figure 6.36: Schematic of Impact Test Setup: (1) high-speed digital camera,
(2) drop tower impactor tup, (3) delay generator, (4) specimen,
(5) light sources, (6) DAQ – drop tower, (7) DAQ – camera,
(8) lamp control unit, and (9) drop tower controller………………… 134

Figure 6.37: Impact Test Setup: (1) high-speed camera, (2) drop tower impactor
tup, (3) specimen, and (4) light source ……………………………… 135

xiii
Figure 6.38: Camera Viewing Area………………………………………………... 136

Figure 6.39: ARAMIS™ Results for IPC30 (DYN-29, u)………………………… 137

Figure 6.40: ARAMIS™ Results for IPC30 (DYN-29, εx)………………………… 138

Figure 6.41: ARAMIS™ Results for SFSb30 (DYN-35, u)………………………. 139

Figure 6.42: ARAMIS™ Results for SFSb30 (DYN-35, εx)………………………. 140

Figure 6.43: Strain Energy History, IPC30………………………………………… 141

Figure 6.44: Strain Energy History, SFSb30………………………………………. 142

Figure 7.1: Normalized Measured Peak Load for Sandwich Structures with SF
and IPC Foam Cores…………………………………………………. 147

Figure 7.2: Normalized Measured Mid-Point (Load-Point) Deflection at Failure


for Sandwich Structures with SF and IPC Foam Cores……………… 147

Figure 7.3: Normalized Measured Load at Transition Point for Sandwich


Structures with SF and IPC Foam Cores…………………………….. 148

Figure 7.4: Normalized Measured Mid-Point (Load-Point) Deflection at


Transition Point for Sandwich Structures with SF and IPC Foam
Cores………......................................................................................... 148

Figure 7.5: Energy Absorbed for Sandwich Structures with SF and IPC Foam
Cores ………………………………………………………………… 149

Figure 7.6: Specific Energy Absorbed for Sandwich Structures with SF and IPC
Foam Cores ………………………………………………………….. 149

Figure 7.7: 20% Vf: Strain (εx) Variation Along the Height of the Sandwich
Core………………………………………………………………….. 151

Figure 7.8: 20% Vf: Shear Strain (εxy) Variation Along the Height of the Sandwich
Core …………………………………………………………………. 152

xiv
List of Tables

Table 2.1: Syntactic Foam Constituent Properties………………………………… 17

Table 2.2: Duocel® Aluminum Foam Properties………………………………… 17

Table 3.1: Aluminum Foam Sandwich Test Results …………………………….. 31

Table 3.2: TPB SFS20 Test Results ……………………………………………… 34

Table 3.3: TPB SFS30 Test Results ……………………………………………… 34

Table 3.4: TPB SFS40 Test Results ……………………………………………… 35

Table 3.5: Material Densities……………………………………………………... 41

Table 3.6: Syntactic Foam Quasi-Static Elastic Properties [17]…………………. 43

Table 4.1: TPB IPC20 Test Results ……………………………………………… 51

Table 4.2: TPB IPC30 Test Results ……………………………………………… 51

Table 4.3: TPB IPC40 Test Results ……………………………………………… 52

Table 4.4: IPC Material Densities………………………………………………… 62

Table 4.5: IPC Quasi-Static Elastic Properties [17]………………………………. 66

Table 5.1: TPB SFSb20 Test Results …………………………………………….. 75

Table 5.2: TPB SFSb30 Test Results …………………………………………….. 75

Table 5.3: TPB SFSb40 Test Results …………………………………………….. 76

Table 6.1: IPC30 Impact Test Results …………………………………………… 142

Table 6.2: SFSb30 Impact Test Results …………………………………………. 142

xv
CHAPTER 1

INTRODUCTION

1.1 Cellular Solids and Foams

A cellular solid is made up of an interconnected network of solid struts or thin

plates that form the edges or faces of cells. The simplest of all cellular solids is a two-

dimensional array of regular polygons that look like the cells created by honeybees. For

this reason they are often referred to as honeycombs. More commonly seen are three-

dimensionally packed polyhedra which are known as foams. If the cells connect through

open faces with the cell edges that are solid, the foam is said to be open-celled. On the

other hand, if the cell faces are also solid, sealing off each cell from its neighbors, the

foam is said to be closed-celled [1]. Figures 1.1-1.3 show images of these three foam

types.

Figure 1.1 Honeycomb Foam (Ref:

http://materiali.matech.it/matech/materiali/images/okpic/CP2079.jpg)

1
Figure 1.2 Open-Celled Foam (Ref: http://www.grantadesign.com/images/foam.gif)

Figure 1.3 Closed-Celled Foam (Ref: http://www.alcarbon.de/jcms/images/stories/Al-

Schaeume/ALPORAS_AC/ALPORAS.JPG)

Many structural materials found in nature are cellular solids: cork, sponge, coral,

wood, and cancellous bone. Wood is still the most widely used structural material.

Increasingly, man-made foams and honeycombs are being used in structural applications.

Considering their lightweight and stiff nature, their most obvious use today is in

sandwich panels. From the World War II planes that used panels made from thin

plywood skins bonded to balsa wood cores to the planes of today that use fiberglass or

2
carbon fiber composite skins with cores of aluminum or paper-resin honeycombs or rigid

polymer foams, sandwich panels provide excellent specific bending stiffness and

strength. Sandwich panels are also found in nature. The cuttlebone of a cuttlefish is a

multi-layer sandwich panel. Some types of leaves are structured much in the same

manner as sandwich panels. Even the human skull is composed of two layers of hard

bone separated by a lightweight core of cancellous bone. Figure 1.4 shows the cancellous

bone found in the human skull compared with Duocel® open-celled aluminum foam.

(a) (b)

Figure 1.4 (a) Cancellous Bone and (b) Duocel® Aluminum Foam (Ref:

http://www.ergaerospace.com/foamproperties/aluminumproperties.htm)

The single most important structural characteristic of cellular solids is relative

density, the density of the foam divided by the density of the solid of which it is

comprised.

Syntactic foams (SFs) are a class of structural composite foams created by

dispersing hollow microballoons into a metal, polymer, or ceramic matrix. The presence

3
of the microballoons leads to improved properties including buoyancy, lower density, and

higher strength.

1.2 Sandwich Panels with Foam Cores

Sandwich panels consist of two stiff, strong skins separated by a lightweight core.

Separating the skins, or face sheets, increases the moment of inertia of the panel with

little penalty in terms of weight. This produces an efficient structure for withstanding

bending and buckling loads. For this reason sandwich panels are often used in situations

where weight limitation is critical, such as in aircraft structures. The conventional way of

stiffening panels are with stringers: strips attached between the faces with profiles such as

a Z, a tee, or a top hat. In general, stringer stiffened panels offer greater specific stiffness

(stiffness per unit weight). The drawbacks, however, are anisotropy and cost. Metal or

syntactic foam cores in sandwich structures offer advantages of macroscopic material

isotropy as well as the ability to easily and cheaply mold complex lightweight structures

in a single operation. For applications involving a dominant bending moment,

attachment or honeycomb stiffened sandwich panels can offer the best specific stiffness.

This applies only in the direction the panel has been stiffened, though. In contrast, for

loads about which less information is known or that have comparable magnitudes in

multi-axial (even changing) directions, the isotropic nature of foam cored sandwich

panels offer the greatest specific stiffness in all directions.

4
Figure 1.5 Sandwich Panels with Metallic Foam Core (Ref: http://www.agstaron.com)

1.3 Interpenetrating Phase Composites (IPCs)

Interpenetrating phase composites (IPCs) are a relatively new class of composite

materials that has the potential for outstanding multifunctional properties. In traditional

composites, discrete reinforcing phases such as whiskers, fibers, or dispersed particles are

introduced into a matrix to improve or alter the resulting properties of the matrix. In an

IPC, constituent phases are continuous and three dimensionally interconnected. In their

singular state, each constituent of an IPC would be a matrix consisting of an open-celled

microstructure. Hence, IPCs are uniquely different from traditional composites

consisting of a matrix with reinforcing phases that do not experience such complete

interpenetration. Figure 1.5 illustrates an idealized open-celled nature of two phases of

an IPC.

Figure 1.6 Idealized Interpenetrating Phase Composite Network

5
An advantage of IPCs over traditional reinforced composites is that higher

volume fractions of the reinforcing or secondary phase may be introduced more easily

into an IPC. Furthermore, each phase is mechanically constrained by the other resulting

in a better structural response. Each phase of an IPC contributes its property to the

overall macro scale characteristics synergistically. For example, if one constituent

provides strength, the other might enhance stiffness or thermal stability. Among the

potential mechanical benefits of IPCs, the ones regarding fracture and energy dissipation

characteristics are of interest. In traditional composites with aligned fibers, stiffness and

strength advantages are limited to the fiber direction as crack propagation along the fibers

cannot be effectively resisted. On the contrary, the three dimensional interconnectivity of

the phases in an IPC could stave off failure effectively while offering macro scale

isotropy.

1.4 Literature Review

The work done by Wu et al. [2] illustrates improved mechanical properties and

impact resistance of sandwich structures with graphite/epoxy face sheets and aluminum

honeycomb core by first filling the honeycomb core with rigid polyurethane foam. Low

velocity impact tests have revealed an improvement in impact resistance, including less

frequent debonding of the face sheet and localized core crushing. Zarei et al. [3]

attempted to study and optimize bending behavior due to impact loading of foam-filled

beams. Impact tests were conducted on aluminum tubes in order to simulate car bumpers

6
during crashes. An attempt was made to optimize the tube energy absorption and tube

design to include a metallic foam core consisting of Alporas aluminum foam. Energy

absorption increased between 2.5%-5.2% once the foam filler was added. Specific

energy absorbed also increased; however, it was less pronounced. Seitzberger et al. [4]

examined the effects of foam filling on axially loaded tubular columns made of steel.

Different shapes, such as square, hexagonal, and octagonal, as well as different filling

configurations (single tube with no foam filling, two tubes with foam in between, single

completely filled tube) were tested. For the square, fully filled tubes, experiments

indicated a specific energy absorption improvement of up to 60%. A glass fiber

composite-foam sandwich structure was studied in bending and impact situations by

Belingardi et al. [5]. Bending tests failed starting with 45 degree shear cracks in the core

followed by face sheet debonding. The modulus and Poisson’s ratio gathered from

bending tests were approximately 25 percent below the values from tensile tests, likely

due to the dual tensile and compressive nature of loading. Tests conducted on the

sandwich structure showed no strain-rate effects over the range of test conditions

investigated in the study. The impact response of this foam-based sandwich structure

was successfully modeled using mechanical properties determined at static or quasi-static

rates of strain.

Cenosphere fly ash (waste product from burning coal) microballoons and

aluminum were used to create syntactic foam via pressure infiltration technique by Zhang

et al. [6]. Compression tests revealed energy absorption capacity to be 27 MJ·m-3 at 47%

strain with a peak stress of 73 MPa. Metal-matrix IPCs were created by Zhou et al. [7].

Using volatile agents, an aluminum matrix was created with open porosity as high as

7
83%. The aluminum matrix interpenetrating phase composites reinforced with

continuous Al2O3-TiC ceramic phase were successfully fabricated by the vacuum high

pressure infiltration process.

One work by Styles et al. [8] investigates the strain distribution and failure

mechanisms associated with aluminum foam or polymer foam sandwich structures under

four-point bending. Real-time strain data was collected using a 3D optical technique

(ARAMIS™, GOM mbH, Germany). It was determined that the energy absorbed to

yield load for aluminum foam was approximately two times that of the polymer, and the

total energy absorption of aluminum was about 3.5 times greater. In a separate work by

Styles et al. [9], core thickness effects on the flexural behavior of aluminum foam

sandwich structures were examined. In sandwich material with Alporas foam as the core,

the max core stress and max facing stresses decreased as core thickness increased,

possibly due to diminishing coupling between core and facing. Core shear stress,

however, remained nearly the same (approximately 1.75 MPa). The failure of the core

became more dominant as the core thickness increased. McKown and Mines [10] studied

static and impact behavior of metal foam cored sandwich beams. The sandwich material

consisted of Alporas core with glass fiber polypropylene matrix composite (Plytron) face

sheets. In quasi-static tests, the shear effects in the core increased as the total specimen

length decreased. The range of indenter deflections (deflections=15-30 mm on a

specimen of length=230 mm) for the two test types were similar, indicating that strain

rate effects essentially cancel out the core failure initiation at weak points in the structure.

From indenter energies of 15 J up to 50 J, fraction of energy absorbed up to skin failure

ranged from 45% down to 20%, respectively.

8
The work done by McCormack et al. [11] illustrated the different modes of

sandwich beam failure under quasi-static bending. Among face yielding, face wrinkling,

core yield, indentation, and delamination, the most common failure modes observed were

core yield and face yield. Chen et al. [12] studied the plastic collapse of sandwich beams

with Alporas™ foam cores. Four-point bending allowed the competing failure modes of

face yield, core shear and indentation to be separated physically along the beam. Face

yield occurred between the inner rollers, core shear occurred between the inner and outer

rollers, and indentation was triggered directly beneath the rollers. The collapse

mechanism maps were generated as a tool to predict failure type/mode, and criteria for

each of the failure cases were generated.

Dukhan et al. [13] performed three-point bending tests on aluminum-

polypropylene IPC material. Strength and flexural modulus of the IPC were compared to

that of polypropylene. The IPC was stiffer than both constituents, and increased with

decreasing pore size. When 20 ppi aluminum net was used, an increase in strength of 3-

4% and an increase in modulus of around 33% accompanied the change from

polypropylene to IPC. The effect of volume fraction on alumina/aluminum composites

was investigated by Travitzky [14]. Various volume fractions of aluminum were

interspersed into alumina. For 12% aluminum (closest to the material presented in this

study) the bending strength increased from about 205 to 740 MPa, a 2.5 times increase.

The Young’s modulus increased from about 205 to 330 GPa, an increase of 60%. Chang

et al. [15] utilized a pressureless infiltration system to produce and characterize an

Al(Mg)/Al2O3 interpenetrating composite. Three-point bending tests revealed an increase

in bending strength from approximately 5 MPa to 250 MPa once a metallic phase was

9
introduced to a foam of 15% relative density. Also, a decreasing trend in IPC strength

was observed as relative density increased.

1.5 Sandwich Beam Theory

This section puts forth the theory associated with bending of sandwich structures

applicable to linear elastic cases. It will be shown in later chapters that both IPC foam

core sandwiches and syntactic foam core sandwiches with graded face sheets (see

Chapter 2 for material descriptions) may be initially linear elastic, but does not remain so

for the entirety of its tests. This section is comprised of such theory associated with

sandwich beams under three-point bending (TPB) loads.

Equations based on classical beam theory are provided by Allen [16] and are

given below for sandwich structures. Consider a sandwich beam subjected to three-point

bending as shown in Figure 1.7.

core face sheet

y y
z
x

Figure 1.7 Three-Point Bending Schematic of a Sandwich Beam

10
Let the flexural rigidity be denoted by EI. For stacked composite beams (e.g.

sandwich structures), the flexural rigidity is the sum of the flexural rigidities of the two

faces and the core. It is generally represented by the symbol D and evaluated as follows:

bc3
Dcore = Ec ,
12

 bt 3 btd 2 
D faces = 2  E f + Ef  ,
 12 4 

bt 3 btd 2 bc3
D = Dcore + D faces = E f + Ef + Ec ,
6 2 12

where the subscripts f and c represent the face sheet and core, respectively. Similarly, the

equivalent shear rigidity, (AG)eq, in stacked composite beams is:

bd 2
( AG ) eq = Gc .
c

where A and G represent the cross section area and shear modulus, respectively. These

rigidities are used to calculate the displacement at the midspan of a (linear elastic) beam

as:

PL3 PL
δ= + .
48 D 4( AG )eq

11
PL3
The first term of the deflection equation, , represents the contribution due to
48 D

PL
bending. The second term, , represents the contribution due to shear.
4( AG )eq

To obtain equations for stress in a three-point symmetric loading configuration,

we begin by recognizing that strain ε=σ/E leads to:

My c h h c
σf =− Ef  ≤ y≤ ; − ≤ y≤−  ,
D 2 2 2 2

My  c c
σc = − Ec − ≤ y ≤  ,
D  2 2

 Px L
 2 , 0≤x≤
2
where moment is given by, M =  ,
 P ( x − L ) ,
L
<x≤L
 2 2

and D represents the flexural rigidity. (Consult Figure 1.7 for definitions of the

geometric parameters.) Ordinary homogeneous beam bending theory gives the

expression for shear stress at depth, y, below the centroid in the cross-section as:

QS
τ= ,
Ib

where Q is the first moment of area, I is the second moment of area about the centroid, S

is the shear force, and b is the beam width. For a compound beam, differing moduli for

different elements in the cross-section must be accounted for using the modification:

S
τ=
Db
∑ (QE ) ,

12
where ∑ (QE ) represents the sum of the products of the first moments of area (Q) and
moduli (E), and D is the flexural rigidity. For the sandwich construction presented here,

btd Ec b  c  c 
∑ (QE ) = E f
2
+
2 2
 + y  − y  .
 2 

Therefore, under three-point bending configuration the shear stress experienced in

the core can be calculated using:

S  td Ec  c 2  
τ= E f +  − y 2  ,
D  2 2  4  

 P L
 2 , 0 ≤ x ≤ 2
where the shear force is given by, S =  .
− P , L < x ≤ L
 2 2

The following are definitions of the symbols used in this and subsequent chapters:

(AG)eq = shear rigidity

b = specimen width

c = core thickness

δ = deflection at mid-span

d = distance between facing centroids

D = flexural rigidity

Ec = modulus of the core

Ef = modulus of the facings

Gc = shear modulus of the core

h = total specimen height

13
l = total specimen length

L = support span

P = applied load

ρrel = relative density of foam

σf = facing stress

σc = core normal stress

τ = core shear stress

t = facing thickness

1.6 Objectives

This work is aimed towards demonstrating the feasibility of a lightweight

sandwich structure with epoxy-based syntactic foam (SF) core and graded face sheets and

an SF-based IPC foam core sandwich for improving the flexural response relative to that

of a similar sandwich with a simple SF core. The specific goals for the work are:

• Develop a method for processing sandwich structures with syntactic foam and

IPC foam cores.

• Obtain both global and local flexural response data of SF and IPC core sandwich

structures.

• Explain failure mechanisms under quasi-static three-point bending configuration.

• Study the effect of hollow filler volume fraction in the syntactic foam (20%, 30%,

or 40%) on flexural characteristics.

14
• Study the possibility of improving load-deflection and strain energy absorption

responses of syntactic foam core sandwiches by either (a) adding a 3D

interconnected metallic foam phase to the core or (b) tailoring the face sheets to

prevent face-core debonding.

• For both quasi-static and impact loading cases, demonstrate ability to make local

displacement and strain measurements using digital image correlation method in

sandwich structures.

1.7 Organization of the Thesis

This thesis is divided into seven chapters, including this one. The first chapter

describes the materials to be investigated, motivation for the research, and previous work

by others in related areas. Material description and preparation of the syntactic foam core

and IPC foam core sandwiches are discussed in Chapter 2. The flexural characterizations

of unfilled aluminum foam core sandwiches and syntactic foam core sandwiches are

discussed in Chapter 3. Chapter 4 contains discussion of flexural characterization of IPC

foam core sandwiches. Chapter 5 contains discussion of flexural characterization of

syntactic foam core sandwiches with graded face sheets. Chapter 6 is concerned with the

digital image correlation method used to analyze full-field deformations. Finally Chapter

7 presents a comparative summary and conclusions to be drawn from this work. There

are appendices that show global/local measurement comparisons and images related to

the ARAMIS™ digital image correlation work.

15
CHAPTER 2

MATERIAL DESCRIPTION AND SAMPLE PREPARATION

2.1 Material Description

Sandwich structures with aluminum faces and syntactic foam (SF) or SF-based

IPC foam core are processed in this work. Pressureless infiltration techniques are

employed to create the syntactic foam-filled IPC. The IPC foam is created by

introducing uncured syntactic foam into aluminum sandwich performs and curing the SF

in situ.

2.1.1 Syntactic Foam

The syntactic foam used in this work is made of epoxy matrix and hollow glass

microballoon filler. Physical and elastic properties of the constituents of the syntactic

foam used are shown in Table 2.1. The epoxy used was Epo-Thin®, a low viscosity resin

(supplied by Buehler, Inc., USA). The advantage of the Epo-Thin® epoxy is its low

viscosity, necessary to fill small holes and voids between microballoons. The hollow

glass microballoons (K-1™ microballoons supplied by 3M Corp.) were of average

diameter ~65 µm and wall thickness ~550 nm. The microballoons used are the most

16
economical hollow ceramic fillers produced by 3M Corp. and advantages of their use

include the ability to create a good bond between epoxy and glass and it having one of

the lowest bulk densities among the available glass microballoons.

Table 2.1 Syntactic Foam Constituent Properties

Property Neat Epoxy Microballons


Elastic Modulus (MPa) 3200 -
Bulk Density (kg/m3) 1085 125
Poisson's Ratio 0.34 -

2.1.2 Aluminum Foam

The open cell aluminum foam was used as the scaffold for infusing SF into the

core of the IPC sandwiches. The scaffold material was Duocel® foam obtained from

ERG Materials and Aerospace Corp. Its properties are shown in Table 2.2. Mean cell

size for 20 ppi material is ~1.02mm (0.040”). Ligament diameter is ~0.2mm (0.008”).

Table 2.2 Duocel® Aluminum Foam Properties

Material Pore Density Relative Density


Aluminum
20 ppi 7%
6101

The open-cell scaffold is made from annealed 6101 aluminum, and the face sheets

of the sandwich structures are made from annealed 6061 aluminum. Relevant elastic

properties for these two alloys are the same, Young’s modulus of 69 GPa (10 Msi) and

17
yield strength of 552 MPa (8 ksi). Figure 2.1 shows the open-celled aluminum foam

sandwich scaffold, and Figure 2.2 shows an enlarged view of the open-cell Duocel®

foam.

Figure 2.1 Open-Celled Aluminum Foam Sandwich Scaffold

Figure 2.2 Enlarged View of Duocel® Foam (Ref:

http://www.ergaerospace.com/foamproperties/introduction.htm)

2.1.3 IPC Foam

The IPC core sandwiches were produced by introducing uncured syntactic foam

into the aluminum foam core scaffold and allowing the SF to cure in situ. From the

results and conclusions from Jhaver [17] which demonstrate IPC foam superiority over

epoxy-based syntactic foam in compression, it was thought that using IPC as the core

18
material in the sandwiches would produce an improved response compared to sandwiches

with an SF core.

2.2 Specimen Preparation

2.2.1 Mold Preparation

The specimens used in this work were made using castable two-part thermosetting

polymer. Given the simplicity of geometry, the specimen was cast into a near net shape

with minor finish machining, reducing the processing time required for individual

experiments. To accomplish this, a specimen blank was machined from a stock

aluminum block. A mold cavity was then produced by pouring silicone rubber around

the blank to create a negative of the specimen. Specimens were then created by casting

the material directly into the mold cavity.

The specific rubber product used for creating the mold was PlatSil® 73 Series

from PolyTek Development Corporation. This is a flexible rubber mold material that has

high tear strength and can be vulcanized at room temperature. It has low shrinkage,

resulting in very good dimensional stability. To create the mold, a cardboard barrier was

constructed and placed on a glass substrate. The specimen blank was placed within the

cardboard barrier and the silicone rubber mixture was poured over the blank and allowed

to cure for 30 hours. After curing, the cardboard barrier and the specimen blank were

removed resulting in a cavity for future specimen castings.

19
2.2.2 Syntactic Foam Core Specimen Preparation

Epoxy-based syntactic foams with varying volume fractions (20%, 30%, and

40%) of hollow soda-lime glass microballoons were produced. For a given test

specimen, the general preparation sequence was as follows. First, the desired quantities

of resin, hardener, and glass particles were measured out. Next, the resin and hardener

were mixed slowly until the mixture appeared homogeneous. Upon achieving visible

homogeneity, the glass particles were added to the mixture and stirred slowly until the

mixture was once again homogeneous. Once the mixture was visibly homogeneous, it

was then placed into a vacuum chamber. A vacuum pump was used to bring the chamber

down to a pressure of approximately -80 kPa (gage). This pressure was maintained for

approximately 4-5 minutes and then released, returning the mixture to atmospheric

pressure. This process of vacuuming and releasing was repeated in order to bring trapped

air to the surface (typically 3 times). This ensured full degassing of the mixture. After

the vacuum/release process, air bubbles were skimmed from the top of the mixing

container. During degassing, the mixture began to separate slightly into epoxy and glass

particle constituents. It was briefly and slowly stirred once more until fully mixed before

pouring into the mold. A photograph of the mold is shown in Figure 2.3 along with a

demolded IPC foam core specimen (to be described later on).

20
Figure 2.3 Silicone Specimen Mold and Demolded Specimen

All castings for this work were allowed to cure at room temperature for 48 hours

prior to removal from the mold. Upon completion of curing, the specimens were finish

machined using a mill. Machining was done no less than 2 days after initial casting, and

testing occurred no less than 7 days after the initial casting.

To produce the Syntactic Foam Sandwich (SFS) specimens, face sheets were

punched out of a 0.813mm thick sheet of aluminum (AL6061) and machined to precise

size. They were sanded on one side with coarse grit sandpaper then cleaned off with

acetone in order to create a rougher surface for bonding. Once the core of syntactic foam

was fully cured and machined, each face sheet was separately adhered to the core using a

syntactic foam mixture of the same volume fraction. After curing, the specimen was

once again machined to remove excess material from the face sheet application process.

Figure 2.4 shows an example of the SFS sandwich structure. Though there is some minor

variation, the nominal specimen size for quasi-static testing was 127mm x 25mm x 20mm

(5” length x 1” width x 0.8” height). Each face sheet had a nominal thickness of

21
0.813mm, giving a core thickness of ~18.4mm. The nominal specimen size used for

dynamic testing was 127mm x 25mm x 18.5mm (5” length x 1” width x 0.73” height).

0.813mm thk

SF core

Aluminum
face sheet

Figure 2.4 Syntactic Foam Core Sandwich Structure (SFS)

2.2.3 IPC Foam Core Specimen Preparation

The open cell aluminum foam sandwich material was used as the IPC scaffold.

The foam was cut into slightly oversized pieces compared to desired final specimen size.

The face sheet thickness for all specimens was in an as-received state of 0.813mm ±

0.05mm (0.032” ± 0.002”). The aluminum was first cleaned with acetone, then coated

with silane to act as a wetting agent and to promote adhesion between the aluminum

scaffold and syntactic foam. Jhaver [17] explored and confirmed the benefit of the silane

coating prior to introducing the syntactic foam. The silane used was A-1100 Amino

Silane, or γ-aminopropyltriethoxysilane (H2NCH2CH2CH2Si(OCH2CH3)3), obtained from

GE Silicones. A 5% silane solution was created (remaining 95% portion of mixture was

9 parts ethanol to 1 part deionized, filtered water) to coat the sample per the work of

22
Sadler and Vecere [18]. The solution was allowed to rest for one hour after mixing

before being used. After the coating, the aluminum was allowed to sit and dry for 12

hours. The exterior/outside of the face sheets were coated with a releasing agent to easily

be able to remove excess syntactic foam after curing. The syntactic foam was prepared in

the same manner as outlined in Section 2.2.2 and poured into the mold, and the aluminum

scaffold was gently pushed down into the pool of uncured syntactic foam.

As before, all castings for this work were allowed to cure at room temperature for

48 hours prior to removal from the mold. Upon completion of curing, the specimens

were finish machined using a mill. Testing occurred no less than 7 days after the initial

casting. A sample of the IPC foam core sandwich structure is shown in Figure 2.5. Note

that the regular diagonal marks easily observed on the aluminum are tool marks.

Syntactic foam

Aluminum foam
ligaments

AL6061
face sheet

Figure 2.5 IPC Foam Core Sandwich Structure

23
2.2.4 Syntactic Foam Core with Graded Face Sheets Specimen Preparation

The graded face sheets (see Section 4.3 and Chapter 5) were obtained from the

same supplier as the aluminum foam core sandwiches. The face sheets had a ~2 mm

layer of aluminum foam brazed to the aluminum face sheets on one side. The face sheets

were cleaned with acetone, then coated with silane in the method described in the

previous section. The exterior/outside of the face sheets were coated with a mold

releasing agent, then placed into the rubber mold as shown in Figure 2.6. The syntactic

foam was prepared, poured into the mold, and allowed to cure in situ. Figure 2.7 shows

an image of the syntactic foam core sandwich with graded face sheets (SFS-b). Note that

the regular diagonal marks easily observed on the face sheets are tool marks.

1 2 3

Figure 2.6 Single Cast SFS-b Mold: (1) Mold, (2) AL6061 Face Sheet, and (3) Brazed

Aluminum Foam

24
Aluminum foam

SF core

AL6061
face sheet

} ”graded” region
Figure 2.7 Syntactic Foam Core Sandwich Structure with Graded Face sheets (SFS-b)

2.2.5 Pattern Preparation for Optical Measurement

In order to perform local displacement and strain measurements, a stochastic

speckle pattern was created on one side of the specimen. The side to which the pattern

was applied was sanded using a fine grit paper and cleaned off using acetone. Once the

specimen surface was clean, alternating coats of white and black spray paint were applied

until a suitable pattern was created (approximately 4 thin coats each). After applying the

speckle pattern, the paint was allowed to dry for 24 hours before testing. Figure 2.8

shows an image of a specimen with the speckle pattern used for obtaining optical

measurements. Approximate speckle size was ~90 µm. Note that the larger black

lines/dots on the specimen are made intentionally for spatial and scaling purposes.

25
loading point

line of symmetry

x mm right support

Figure 2.8 Typical Decorated Random Speckle Pattern for Optical Measurements

26
CHAPTER 3

FLEXURAL CHARACTERISTICS OF CONSTITUENTS

3.1 Aluminum Foam Sandwich

Unfilled aluminum foam sandwich beams were tested in quasi-static three-point

bending (TPB). Although the basis of the TPB tests was ASTM C393 [19], the specimen

size guidelines could not be strictly adhered due to material availability and cost. Instead,

the test specimens resemble short beams. Again, the nominal specimen size for all quasi-

static test specimens was 127mm x 25mm x 20mm (5” length x 1” width x 0.8” height).

An MTS QTest100 (10kN load cell) universal loading machine was used for

flexural tests. A cylindrical loading tip of diameter 10 mm was driven downward at a

velocity of 0.025 in/min (0.635 mm/min) as load-displacement data was collected at a

sampling rate of 5-10 points per second. The beams were centrally positioned on roller

supports, also of diameter 10 mm. The test arrangement is shown in Figure 3.1.

27
CROSSHEAD BEAM

LIGHT LIGHT
SOURCE SOURCE

LOADING
FIXTURE

CAMERA
LENS

Figure 3.1 Photograph of Three-Point Bending Test Setup

In order to collect gray scale data of the speckle pattern for an open-cell foam, the

pores had to be filled with a material that offers negligible reinforcement. Common

expanding insulating foam was used to cover one side (face) of the foam, creating a more

solid and even surface on which speckles could be painted. After the foam had time to

cure, it was sanded down to the surface of the metal foam. The speckle pattern was then

applied as detailed in Section 2.2.4. It was determined through a series of tests that the

insulating foam would not affect the material behavior because its load bearing or

reinforcement capacity is negligible.

28
For quasi-static tests, a camera was set up to capture the right half of the

specimen, between the loading and support rollers. Due to the symmetry of the beam

about the loading axis, one half beam images were captured for subsequent deformation

analysis. Figure 3.2 illustrates the area captured by the camera.

Figure 3.2 Camera Viewing Area

Since a 2D analysis was performed to determine planar surface displacements,

only one digital camera was required to capture the specimen deformation during the

tests. The camera used for quasi-static testing was a Nikon D100 SLR digital camera

fitted with a 200mm lens. Its maximum image capture size is 3008 x 2000 pixels, but for

this work a spatial resolution of 1504 x 1000 pixels was chosen. The camera was

connected to and controlled by a computer using the Nikon Capture software (Version

4.0). The camera was set approximately two feet from the specimen, and white light

illumination sources were placed next to the specimen. The response of the material was

expected to be symmetric about the loading roller; therefore the images encompass the

right half of the beam only, from just outside the loading roller to just outside the support

roller. The camera was set to automatically record images once every 2-3 seconds and

store them on the computer for later analysis. The arrangement for the quasi-static tests

is shown below in Figure 3.3.

29
TPB test digital controls camera & stores
camera images

Illumination
sources

Figure 3.3 Quasi-static Test Experimental Arrangement

The results from the tests on the unfilled aluminum foam core sandwiches are

shown in Figure 3.4 and summarized in Table 3.1.

600 Material: ALS


Temperature: 70°F
Overlay

500

400
FLEX-32 (ALS)
Load (N)

300 FLEX-33 (ALS)

FLEX-34 (ALS)

200

100 **Note that the dotted black


line is the estimated
behavior of FLEX-32 and
not measured data. The
solid black line is measured
data.
0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5
Crosshead Displacement (mm)

Figure 3.4 Overlay of Load-Deflection Data for 20ppi Aluminum Foam Core Sandwich

30
Table 3.1 Aluminum Foam Sandwich Test Results

Crosshead
Specimen Speed Plateau Peak
Number Type (in/min) Load (N) Load (N)
32 ALS 0.025 391 395
33 ALS 0.10 342 346
34 ALS 0.10 391 399
NUM OF SAMPLES 3 3
AVERAGE 375 380
STANDARD DEVIATION 28 30

The unfilled aluminum foam sandwiches behave in a linear fashion at the very

beginning before the onset of nonlinearity due to core indentation and crushing. The load

remains nearly constant throughout the core crushing at the so-called “plateau load” until

the core is free of pores and densification begins (not seen in the displacement range

used). Figure 3.5 shows a typical stress-strain response for aluminum foam and

illustrates the plateau and densification regions. For the case of a TPB test, these

crushing and densification phenomena are observed below the load roller. When in the

plateau region, the aluminum foam has tremendous energy absorption capacity,

continuing to consume energy with increasing strain (or deflection) at a nearly constant

load. The average plateau load for the aluminum foam core sandwich is 375 N ± 28 N.

31
Plateau Region (Core Crushing) Densification Region

Figure 3.5 Typical Aluminum Foam Stress-Strain Response (courtesy: ERG)

3.2 Syntactic Foam Sandwich

The quasi-static TPB test setup for the syntactic foam core sandwich (identified as

SFS) specimens is the same as that detailed in Section 3.1. In order to remove any slack

from the loading fixtures, a pre-load of ~130 N was applied. After this pre-load was

applied, the load was zeroed so that the pre-loaded state is the reference (undeformed)

state. With failure on the order of 5,000 N, the pre-load amount considered was

negligible.

The measured load-deflection responses from the tests are shown in Figures 3.6 to

3.8, and the results are summarized in Tables 3.2 to 3.4. Note that the stair-case pattern

observed in the test data is attributed to the data acquisition rate and software and not to

the mechanical response of the sandwich structure. The different materials with 20%,

32
30%, and 40% volume fractions of microballoons in the SF core will henceforth be

referred to as SFS20, SFS30, and SFS40, respectively.

9000 Material: SFS20


Temperature: 70°F
Overlay

8000

7000

6000
FLEX-37 (SFS20)

FLEX-38 (SFS20)
Load (N)

5000
FLEX-45 (SFS20)

FLEX-46 (SFS20)
4000
FLEX-47 (SFS20)

FLEX-48 (SFS20)
3000

2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Crosshead Displacement (mm)

Figure 3.6 Load-Deflection Data for SFS20

9000 Material: SFS30


Temperature: 70°F
Overlay

8000

7000

6000
FLEX-39 (SFS30)
Load (N)

5000 FLEX-40 (SFS30)

FLEX-41 (SFS30)
4000
FLEX-42 (SFS30)

FLEX-43 (SFS30)
3000
FLEX-44 (SFS30)

2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Crosshead Displacement (mm)

Figure 3.7 Load-Deflection Data for SFS30

33
9000 Material: SFS40
Temperature: 70°F
Overlay

8000

7000

6000

FLEX-23 (SFS40)
Load (N)
5000
FLEX-24 (SFS40)
4000
FLEX-36 (SFS40)

3000

2000

1000
Debond initiation

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Crosshead Displacement (mm)

Figure 3.8 Load-Deflection Data for SFS40

Table 3.2 TPB SFS20 Test Results

Crosshead
Specimen Speed Energy Specific Energy Peak Deflection
Number Type VF(%) (in/min) Absorbed (J) Absorbed (J/kg) Load (N) at Failure (mm)
37 SFS 20 0.025 4.78 68.5 6626 1.24
38 SFS 20 0.025 2.17 31.4 4997 0.81
45 SFS 20 0.025 5.92 84.3 7098 1.42
46 SFS 20 0.025 6.18 88.4 7155 1.45
47 SFS 20 0.025 1.72 24.9 4422 0.71
48 SFS 20 0.025 7.57 108.6 7568 1.68
NUM OF SAMPLES 6 6 6 6
AVERAGE 4.73 67.7 6311 1.22
STANDARD DEVIATION 2.33 33.2 1289 0.38

Table 3.3 TPB SFS30 Test Results

Crosshead
Specimen Speed Energy Specific Energy Peak Deflection
Number Type VF(%) (in/min) Absorbed (J) Absorbed (J/kg) Load (N) at Failure (mm)
39 SFS 30 0.025 3.53 57.1 5526 1.24
40 SFS 30 0.025 3.47 55.7 5550 1.17
41 SFS 30 0.025 4.14 65.9 5815 1.37
42 SFS 30 0.025 4.08 64.9 5758 1.27
43 SFS 30 0.025 3.29 52.5 5405 1.14
44 SFS 30 0.025 3.57 57.5 5571 1.22
NUM OF SAMPLES 6 6 6 6
AVERAGE 3.68 58.9 5604 1.24
STANDARD DEVIATION 0.35 5.3 154 0.08

34
Table 3.4 TPB SFS40 Test Results

Crosshead
Specimen Speed Energy Specific Energy Peak Deflection
Number Type VF(%) (in/min) Absorbed (J) Absorbed (J/kg) Load (N) at Failure (mm)
36 SFS 40 0.025 2.78 48.5 5126 0.97
23 SFS 40 0.025 2.35 40.5 3598 1.24
24 SFS 40 0.025 1.76 30.6 5440 0.76
NUM OF SAMPLES 3 3 3 3
AVERAGE 2.30 39.9 4722 0.99
STANDARD DEVIATION 0.51 9.0 985 0.24

The flexural response of all SFS sandwiches showed a linear elastic behavior

initially. Failure in the SFS sandwiches typically manifested in a face-core debond.

Once this debond occurs, the majority of load is carried by the core causing it to fail soon

thereafter. To give a visual of the debond failure, Figure 3.9 shows an image captured

soon after the debond occurred during a test. (Analyzed digital images of this specimen

(FLEX-30, SFS30) are shown in Section 6.4.1.)

Face-core debonding

Figure 3.9 Typical SFS Face-Core Debond

35
The metrics used to characterize the syntactic foam core sandwiches were the

measured peak load, the deflection measured at failure, the total strain energy absorbed,

and the specific strain energy absorbed (which takes into account the weight difference

associated with the different volume fractions tested). Both the peak load and deflection

at failure were normalized by values from the behavior of unfilled aluminum foam

sandwiches.

The average value for the peak load decreases with increasing volume fraction.

The average for SFS20 is 6,311 N, for SFS30 is 5,604 N, and for SFS40 is 4,722 N.

Ideally, data would be normalized by a sandwich consisting of aluminum face sheets

separated by an empty core. Practically, the unfilled 20ppi aluminum foam core

sandwich was used to normalize the data. To compare to the behavior of the aluminum

foam core sandwich, the average load at failure for the SFS sandwich is normalized by

the plateau load experienced by the aluminum foam core sandwich (390 N). The average

normalized load at failure for SFS20 is 16.14, for SFS30 is 14.33, and for SFS40 is 12.08.

In Figure 3.10 these values are shown as a histogram. This comparison is not exact. The

plateau load for the aluminum foam core sandwich is the load at which the stress remains

nearly constant while able to withstand a large increase in strain. It is this load that is

typically used as the desired design load. (See Figure 3.5 for a typical stress-strain

response of this type.) The SFS sandwiches do not experience a similar near constant

load, but rather a steady increase (hardening behavior) until failure. This load at failure,

when compared to the plateau load, is twelve to sixteen times higher than the design load

for unfilled aluminum foam core sandwiches. Both the plateau load for the unfilled

36
aluminum foam core and the peak load for the SFS sandwiches can be used as the design

load, so, while inexact, the comparison is justified. The SFS sandwiches show superior

load bearing capacity compared to their unfilled aluminum foam counterpart. The large

error seen in the SFS20 and SFS40 samples is caused by specimens that prematurely

failed to due an especially weak face-core bond, resulting in lowering all reported

average values.

25

SFS

20
Peak Load/ALS Plateau Load

15

10

0
20% 30% 40%
Volume Fraction

Figure 3.10 Normalized Measured Peak Load for Sandwich Structures with SF Foam

Core with Different Volume Fractions

The average value for the mid-span deflection at failure decreases from SFS30 to

SFS20 to SFS40, though there is no distinguishable difference in SFS20 and SFS30.

However, without the two poorly bonded specimens in the SFS20 group that failed

relatively earlier, the trend would have again decreased with increasing volume fraction.

37
The average for SFS20 is 1.22 mm, for SFS30 is 1.24 mm, and for SFS40 is 0.99 mm.

To compare with the behavior of the aluminum foam core sandwich, the average

deflection at failure for the SFS sandwich is normalized by the deflection at which the

unfilled aluminum foam core sandwich behavior deviates from linearity (deflection at

transition, 0.2mm). This, again, is not an exact comparison since we do not have a

deflection at failure for comparison in the unfilled aluminum foam sandwiches. The

average normalized deflection at failure for SFS20 is 6.29, for SFS30 is 6.38, and for

SFS40 is 5.11. In Figure 3.11 these normalized deflection values are shown as a

histogram. Once again, the large error in the SFS20 and SFS40 samples is due to

specimens that experienced premature failure.

9.0
SFS
8.0

7.0

6.0
δFailure/δALS, nonlinear

5.0

4.0

3.0

2.0

1.0

0.0
20% 30% 40%
Volume Fraction

Figure 3.11 Normalized Measured Mid-Span (Load-Point) Deflection at Failure for

Sandwich Structures with SF Foam Core

38
Cellular foams are used in packing and automotive applications because of their

ability to dissipate strain energy. Using this lightweight sandwich material could

hopefully offer improved mechanical characteristics while still maintaining a low weight.

The energy absorbed during flexural experiments by a specimen was calculated using the

area under the load-deflection curve from the measured data shown in Figures 3.6 to 3.8.

That is, the area below the load-deflection responses were computed as,

δ
U = ∫ P (δ ) d δ ,
0

by performing numerical integration (trapezoid method) of the measured data.

The average values for the strain energy absorbed by different SFS samples are

plotted in Figure 3.12 as a histogram. The average for SFS20 is 4.73 J, for SFS30 is 3.68

J, and for SFS40 is 2.30 J. Unfilled aluminum foam core sandwiches have a great

capacity for energy absorption, but the cost is that they are unable to bear very much

load. With these SFS sandwiches, an effort was made to understand how much energy

could be absorbed in a material with good strength bearing abilities.

39
7.0

SFS
6.0

5.0

Energy Absorbed (J)


4.0

3.0

2.0

1.0

0.0
20% 30% 40%
Volume Fraction

Figure 3.12 Energy Absorbed for Sandwich Structures with SF Foam Core

In order to make comparisons for varying volume fractions (and thus the

difference in weight), specific energy (J/kg) was calculated for each specimen. The

energy absorbed was divided by the mass of the particular specimen determined by the

equation,

mtot = ρcoreVcore + ρ facesV faces .

The densities for the SF core and aluminum faces are shown in Table 3.5.

40
Table 3.5 Material Densities

Density
Material
(kg/m3)
SF20 893.0
SF30 797.0
SF40 701.0
Al 6061 (face) 2700.0

Once again we see the trend of the average value for the specific energy absorbed

decreasing with increasing volume fraction. The average for SFS20 is 67.7 J/kg, for

SFS30 is 58.9 J/kg, and for SFS40 is 39.9 J/kg. Figure 3.13 shows these values as a

histogram. Most likely it is specific energy that applications involving typical

lightweight sandwich structures should use as a primary design metric.

120

SFS

100
Specific Energy Absorbed (J/kg)

80

60

40

20

0
20% 30% 40%
Volume Fraction

Figure 3.13 Specific Energy Absorbed for Sandwich Structures with SF Foam Core

41
As stated earlier, the SFS sandwich failures manifested in face-core debonding.

Figure 3.14 gives a visual of a failed SFS sandwich. Note the lack of any residual SF

remaining on the interior portions of the face sheets in image (b), highlighting the poor

bond strength and ease of separation between the core and face.

(a) (b)

Figure 3.14 (a) Failed Syntactic Foam Core Sandwich with (b) Highlighted Face-core

Debond

The overall behavior of SFS sandwich structure is essentially linear elastic almost

until failure. The response of a typical specimen at each volume fraction was compared

to the theoretical calculations from Section 1.5.1 [16]. The load histories are used to

calculate the theoretical displacements using geometric and elastic parameters. To

theoretically calculate linear elastic mid-span displacements, the quasi-static elastic

properties used for the SF material measured by Jhaver [17] under compression were

derived from the stress-strain responses shown in Figure 3.15. The values are shown in

Table 3.6. (This was an approximation made deliberately even though sandwich beams

experience both compressive and tensile stresses, and tensile stress-strain responses could

42
be different from the compression responses.) There is good agreement between the

theoretical and experimental mid-span displacements for all volume fractions, seen in the

comparisons shown in Figure 3.16 to 3.18. Finally, Figure 3.19 shows a load-

displacement overlay of the three specimens used as examples, with each containing a

different volume fraction of SF in the core.

Figure 3.15 Compressive Response of Syntactic Foam [17]

Table 3.6 Syntactic Foam Quasi-Static Elastic Properties [17]

Elastic Modulus
Material
(MPa)
SF20 1595
SF30 1448
SF40 1261

43
Material: SFS20
6000 Temperature: 70°F
Specimen: FLEX-38
Cast: 12Jul2010
Test: 14Sept2010
Rate: 0.025 in/min
5000

4000
Load (N)

Crosshead
3000

Elastic
Prediction
2000

1000

0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Displacement (mm)

Figure 3.16 Predicted vs. Experimental Mid-Span Deflection, SFS20

Material: SFS30
6000 Temperature: 70°F
Specimen: FLEX-39
Cast: 04Oct2010
Test: 26Oct2010
Rate: 0.025 in/min
5000

4000
Load (N)

Crosshead
3000

Elastic
Prediction
2000

1000

0
0.0 0.2 0.4 0.6 0.8 1.0 1.2
Displacement (mm)

Figure 3.17 Predicted vs. Experimental Mid-Span Deflection, SFS30

44
Material: SFS40
4500 Temperature: 70°F
Specimen: FLEX-24
Cast: 18May2010
4000 Test: 15Jul2010
Rate: 0.025 in/min

3500

3000
Load (N)

2500
Crosshead

2000
Elastic
Prediction
1500

1000

500

0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Displacement (mm)

Figure 3.18 Predicted vs. Experimental Mid-Span Deflection, SFS40

9000 Material: SFS


Temperature: 70°F
Overlay

8000

7000 20%

6000 FLEX-38 (SFS20)


SFS30
Load (N)

5000 SFS20 30%

4000 FLEX-39 (SFS30)

40%
3000

FLEX-24 (SFS40)
2000 SFS40

1000

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Crosshead Displacement (mm)

Figure 3.19 Overlay of Measured Load-Displacement of Sandwich Structures with

Syntactic Foam Core and Different Volume Fraction of Microballoons in the SF

45
3.3 Summary

It is difficult to speak much to the flexural characteristics of the syntactic foam

core sandwiches. Failure in the SFS sandwiches typically manifested in face-core

debonding, shown in Figure 3.20 where part of the top face has separated from the core.

The separation is clean, with no SF residue on the debonded portion of the face sheet.

Figure 3.20 Face-Core Debond

Once this debond occurs, the majority of load is transferred to the core causing it

to fail soon thereafter. The idea of the sandwich structure is for both the core and faces to

contribute to the overall behavior. After the debond, the material no longer acted as a

lightweight sandwich composite, but rather a syntactic foam beam with aluminum plates

on either side. This suggests that overcoming the shear stress jump at the face-core

interface could prevent face-core debonding. Perhaps the addition of a metal foam matrix

to the core would serve to prevent debonding as well as improve the deflections (and thus

strains) experienced through the synergistic nature of an IPC foam core. This would also

46
serve to increase the energy absorbing capacity. The metal ligaments could serve as a

bridging agent, keeping the integrity of the sandwich structure longer.

47
CHAPTER 4

FLEXURAL CHARACTERISTICS OF IPC FOAM CORE

SANDWICH STRUCTURES

4.1 Experimental Setup

The next step was to characterize sandwich structures with aluminum-syntactic

foam interpenetrating core architecture. It was thought that the introduction of the

metallic foam scaffold into the core would improve the flexural response synergistically,

retaining the strength of the syntactic foam and gaining ductility from the aluminum

foam. The experimental characterization of these structures in bending are carried out

using the same setup described for the SFS sandwiches in Chapter 3. In this case, the

sandwich scaffold was an unfilled aluminum foam sandwich into which uncured

syntactic foam was introduced and allowed to cure in situ. An image of an IPC foam

core sandwich structure is shown in Figure 4.1.

48
Syntactic foam
Aluminum foam
ligaments

AL6061
face sheet

Figure 4.1 IPC Foam Core Sandwich Structure

4.2 Quasi-Static Tests

The load-deflection responses measured for aluminum-syntactic foam

interpenetrating core sandwich (identified as IPC) structures are shown in Figures 4.2 to

4.4, and the results compiled from these tests are summarized in Tables 4.1 to 4.3. The

different materials with 20%, 30%, and 40% volume fractions of microballoons in the

IPC core will henceforth be referred to as IPC20, IPC30, and IPC40, respectively.

49
9000 Material: IPC20
Temperature: 70°F
Overlay

8000

7000

6000
FLEX-9 (IPC20)
Load (N)
5000
FLEX-10 (IPC20)

4000
FLEX-11 (IPC20)

3000

2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Crosshead Displacement (mm)

Figure 4.2 Overlay of Measured Load-Deflection Data for Three Different IPC Core

Sandwich Beams (IPC20)

9000 Material: IPC30


Temperature: 70°F
Overlay

8000

7000

6000
FLEX-12 (IPC30)
Load (N)

5000
FLEX-13 (IPC30)

4000
FLEX-16 (IPC30)

3000

2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Crosshead Displacement (mm)

Figure 4.3 Overlay of Measured Load-Deflection Data for Three Different IPC Core

Sandwich Beams (IPC30)

50
9000 Material: IPC40
Temperature: 70°F
Overlay

8000

7000

6000

FLEX-14 (IPC40)
Load (N)
5000

4000 FLEX-15 (IPC40)

3000

2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Crosshead Displacement (mm)

Figure 4.4 Overlay of Measured Load-Deflection Data for Two Different IPC Core

Sandwich Beams (IPC40)

Table 4.1 TPB IPC20 Test Results

Crosshead
Specimen Speed Energy Specific Energy Peak Deflection Transition Transition
Number Type VF(%) (in/min) Absorbed (J) Absorbed (J/kg) Load (N) at Failure (mm) Load (N) Deflection (mm)
9 IPC 20 0.025 7.34 83.3 5366 1.96 1744 0.30
10 IPC 20 0.025 6.34 80.7 4950 1.83 1472 0.28
11 IPC 20 0.025 7.76 89.7 5880 2.08 1846 0.28
NUM OF SAMPLES 3 3 3 3 3 3
AVERAGE 7.15 84.6 5399 1.96 1687 0.29
STANDARD DEVIATION 0.73 4.6 466 0.13 193 0.01

Table 4.2 TPB IPC30 Test Results

Crosshead
Specimen Speed Energy Specific Energy Peak Deflection Transition Transition
Number Type VF(%) (in/min) Absorbed (J) Absorbed (J/kg) Load (N) at Failure (mm) Load (N) Deflection (mm)
12 IPC 30 0.025 7.46 87.9 5850 1.98 1922 0.30
13 IPC 30 0.025 7.71 93.4 5205 2.18 1557 0.28
16 IPC 30 0.025 8.49 98.9 5682 2.24 1726 0.30
NUM OF SAMPLES 3 3 3 3 3 3
AVERAGE 7.89 93.4 5579 2.13 1735 0.30
STANDARD DEVIATION 0.54 5.5 334 0.13 183 0.01

51
Table 4.3 TPB IPC40 Test Results

Crosshead
Specimen Speed Energy Specific Energy Peak Deflection Transition Transition
Number Type VF(%) (in/min) Absorbed (J) Absorbed (J/kg) Load (N) at Failure (mm) Load (N) Deflection (mm)
14 IPC 40 0.025 5.94 89.6 4445 1.91 1521 0.30
15 IPC 40 0.025 5.11 71.3 4128 1.60 1757 0.25
17 IPC 40 0.025 3.67 64.3 4360 1.70 3861* 1.24*
NUM OF SAMPLES 3 3 3 3 2 2
AVERAGE 4.91 75.1 4311 1.74 1639 0.28
STANDARD DEVIATION 1.15 13.1 164 0.16 167 0.04
*Due to loading/reloading during testing, values may not be indicative of true material behavior

In the case of the SFS sandwiches, the load-deflection behavior was essentially

linear elastic in nature (see Chapter 3). However, for the IPC sandwiches it shows

pronounced nonlinearity. There is a kink or a “knee” in each plot beyond which

noticeable nonlinearity prevails up to failure. This “knee” will be referred to as the

“transition point” in the subsequent discussion. To define the transition point for

analysis, a line was fit approximately through the data before the transition and another

was fit through the data after transition. The point where these two fitted lines intersect is

identified as the transition point. Both the load and deflection at the transition point were

recorded from the graphs. The aluminum foam core sandwiches exhibit a similar trend of

transition from linear to nonlinear behavior. However, unlike with the aluminum foam

core (which has a similar value for the transition deflection), the IPC sandwiches are able

to support additional load while continuing to deform. For design purposes, this

transition point can be treated similar to a yield point. At transition, certain parts of the

structure are undergoing plastic deformation that can never be recovered.

The plots and pictures shown below in Figures 4.5 to 4.9 illustrate the progression

of the test at important points of interest. The first photo is the undeformed specimen

used as a reference image for the particular test. The second photo is at point A (in the

plot in Figure 4.5), the transition point, where the behavior deviates from its initial linear

52
elastic response. The third photo is when a visible crack first appears in the core. There

is a short time after the initial crack that the specimen still has the ability to bear some

reduced amount of load. The fourth photo is at the very end of the test when no load

bearing capacity remains in the sample. It should be noted that although no load bearing

capacity remains, the structure itself is still mostly intact. This is unlike the SFS

sandwiches which tended to fail catastrophically. (Analyzed images of this specimen

(FLEX-10, IPC20) are shown in Section 6.4.2.)

Material: IPC20
6000 Temperature: 70°F
Specimen: FLEX-10
Cast: 30Sept2009
Test: 21Apr2010
Rate: 0.025 in/min
5000

4000
Load (N)

3000
Crosshead

2000
B

1000

A C
0
0.0 0.5 1.0 1.5 2.0 2.5
Displacement (mm)

Figure 4.5 Typical IPC Load-Deflection Curve: Point A – Transition Point, Point B –

Crack Formation, Point C – Loss of Load Bearing Capacity

53
Figure 4.6 Undeformed, P=0 N

Figure 4.7 Point A: Transition Point, P=1165 N

54
Figure 4.8 Point B: Crack Formation, P=3905 N

Figure 4.9 Point C: Loss of Load Bearing Capacity, P=1590 N

55
The same metrics used to characterize the SFS sandwiches were used for IPC

sandwiches (measured peak load, deflection measured at failure, total strain energy

absorbed, and specific strain energy absorbed). In addition, two measured values at the

transition point (load and deflection) were used for the IPC sandwiches to identify the

nonlinear behavior. Both the load and deflection measurements were normalized by

values from the behavior of unfilled aluminum foam sandwiches.

The average value for the peak load is the highest for IPC30, with the ones for

IPC20 and IPC40 being successively lower. The average load at failure for IPC20 is

5,399 N, for IPC30 is 5,579 N, and for IPC40 is 4,311 N. To compare to the behavior of

the unfilled aluminum foam core sandwich, the average load at failure for the IPC core

sandwich is normalized by the plateau load experienced by the aluminum foam core

sandwich (390 N). The average normalized load at failure for IPC20 is 13.81, for IPC30

is 14.27, and for IPC40 is 11.03. In Figure 4.10 these values are shown as a histogram.

The plateau load for the aluminum foam core sandwich is the load at which the stress

remains nearly constant while able to withstand a large increase in strain. It is this load

that is typically used as the desired design load. (See Figure 3.5 for a typical stress-strain

response of this type.) The IPC sandwiches do not experience a similar near constant

load, but rather a steady increase (hardening behavior) until failure. This load at failure,

when compared to the plateau load, is eleven to fourteen times higher than the design

load for unfilled aluminum foam core sandwiches. The IPC sandwiches show superior

load bearing capacity compared to their unfilled aluminum foam counterpart.

56
16
IPC
14

12

Peak Load/ALS Plateau Load


10

0
20% 30% 40%
Volume Fraction

Figure 4.10 Normalized Measured Peak Load for Sandwich Structures with IPC Foam

Core

The average value for the mid-span deflection at failure exhibits the same trend

seen for the peak load in the case of the IPC sandwiches, decreasing from IPC30 to

IPC20 to IPC40. The average for IPC20 is 1.96 mm, for IPC30 is 2.13 mm, and for

IPC40 is 1.74 mm. To compare to the behavior of the aluminum foam core sandwich, the

average deflection at failure for the IPC sandwich is normalized by the deflection at

which the unfilled aluminum foam core sandwich behavior deviates from linearity

(deflection at transition). This, again, is not an exact comparison since we do not have a

deflection at failure for comparison in the unfilled aluminum foam sandwiches. The

average normalized deflection at failure for IPC20 is 10.09, for IPC30 is 11.01, and for

IPC40 is 8.96. In Figure 4.11 these values are shown as a histogram.

57
14.0

IPC
12.0

10.0
δFailure/δALS, nonlinear

8.0

6.0

4.0

2.0

0.0
20% 30% 40%
Volume Fraction

Figure 4.11 Normalized Measured Mid-Span (Load-Point) Deflection at Failure for

Sandwich Structures with IPC Foam Core

The average value for the load at the transition point for IPC20 is 1,687 N, for

IPC30 is 1,735 N, and for IPC40 is 1,639 N. As with the peak load, to compare to the

behavior of the aluminum foam core sandwich, the average transition load for the IPC

core sandwich is normalized by the plateau load experienced by the aluminum foam core

sandwich. The average normalized transition load for IPC20 is 4.32, for IPC30 is 4.44,

and for IPC40 is 4.19. In Figure 4.12 these values are shown as a histogram. This is the

best comparison that can be made between IPC foam and aluminum foam core

sandwiches, as both have a transition point. With the IPC foam core sandwich having a

transition load of around 4 times greater than unfilled aluminum foam sandwiches and a

similar transition deflection, the point at which plasticity begins to set in for the IPC

sandwiches is greatly improved. Also, this observation makes intuitive sense. The IPC

58
core sandwiches experienced a transition load four times higher than aluminum foam

sandwiches with a comparable transition deflection, implying that IPC foam core

sandwiches are stiffer than unfilled aluminum foam core sandwiches, which is known to

be true. Evidently, there is no distinguishable difference in the transition load values

within the experimental scatter, although IPC30 specimens show a slightly higher value

relative to the two others. This suggests that microcrack formation in the tensile region

of the core is likely responsible for the early transition point in IPC foam core

sandwiches.

IPC

5
Transition Load/ALS Plateau Load

0
20% 30% 40%
Volume Fraction

Figure 4.12 Normalized Measured Load at Transition Point for Sandwich Structures

with IPC Foam Core

The average for mid-span deflection at the transition point for IPC20 is 0.29 mm,

for IPC30 is 0.30 mm, and for IPC40 is 0.28 mm. To compare with the behavior of the

59
aluminum foam core sandwich, the average transition deflection for the IPC core

sandwich is normalized by the deflection at which the aluminum foam core sandwich

behavior deviates from linearity (0.2mm). The average normalized transition deflection

for IPC20 is 1.49, for IPC30 is 1.53, and for IPC40 is 1.44. In Figure 4.13 these values

are shown as a histogram. As in the transition load plots, these are essentially the same

for all volume fractions of the SF in the IPC core, showing a ~50% improvement over

aluminum foam core sandwiches. With the increased load bearing capacity, higher

transition load, and a small improvement in transition deflection, the IPC foam core

sandwiches are not only similarly stiff, but much stronger than the unfilled aluminum

foam sandwich counterpart.

1.8
IPC
1.6

1.4

1.2
δTransition/δALS, nonlinear

1.0

0.8

0.6

0.4

0.2

0.0
20% 30% 40%
Volume Fraction

Figure 4.13 Normalized Measured Mid-Span (Load-Point) Deflection at Transition

Point for Sandwich Structures with IPC Foam Core

60
Briefly comparing the IPC foam core sandwiches to the SFS sandwiches, the IPC

seems to be the better core material. The normalized peak load showed a small reduction

of 14.5% for IPC20, 0.4% for IPC30, and 8.7% for IPC40. The deflection at failure,

though, showed a pronounced increase. The normalized deflection at failure showed an

improvement of 60.4% for IPC20, 72.6% for IPC30, and 75.2% for IPC40.

As before, the strain energy (U) absorbed by a specimen was calculated using the

area under the load-deflection curve from the measured data. That is,

δ
U = ∫ P (δ ) d δ .
0

The average values for the strain energy absorbed are plotted in Figure 4.14 as

histograms. The average for IPC20 is 7.15 J, for IPC30 is 7.89 J, and for IPC40 is 4.91 J.

9.0
IPC
8.0

7.0

6.0
Energy Absorbed (J)

5.0

4.0

3.0

2.0

1.0

0.0
20% 30% 40%
Volume Fraction

Figure 4.14 Energy Absorbed for Sandwich Structures with IPC Foam Core

61
In order to make comparisons for varying volume fractions, specific energy (J/kg)

was calculated for each specimen. The strain energy absorbed was divided by the mass

of the particular specimen. The mass was determined by

mtot = ρcoreVcore + ρ facesV faces .

The density for the IPC core was determined using a 93% syntactic foam / 7%

aluminum foam composition. The density of the IPC foam core for a particular volume

fraction of syntactic foam therefore was

ρ IPC = 0.93ρ SF + 0.07 ρ Al 6101 .

The densities for the IPC core are shown in Table 4.4. (Refer back to Table 3.5

for densities for the SF core and faces.)

Table 4.4 IPC Material Densities

Density
Material
(kg/m3)
IPC20 1019.5
IPC30 930.2
IPC40 840.9

The results for specific strain energy are shown in Figure 4.15. Again the same

trend is observed in the IPC sandwiches when comparing specific energy absorption.

62
The average value for the specific energy absorbed is highest for the 30% volume

fraction. The IPC20 and IPC40 material values are both less than the IPC30, indicating

an optimum value for the IPC core somewhere around 30% volume fraction of

microballoons in SF. The average for IPC20 is 84.6 J/kg, for IPC30 is 93.4 J/kg, and for

IPC40 is 75.1 J/kg. Comparing the specific strain energies of IPC foam and syntactic

foam core sandwiches (which accounts for the weight difference in adding the aluminum

foam in the core), IPC20 had a 25% improvement over its syntactic foam core

counterpart, IPC30 had a 59% improvement, and IPC40 had an 88% improvement. With

IPC foam core outperforming syntactic foam core for all volume fractions, it can be

concluded that IPC foam core sandwiches are superior to a similar configuration with

syntactic foam core in their energy absorbing capabilities.

120

IPC

100
Specific Energy Absorbed (J/kg)

80

60

40

20

0
20% 30% 40%
Volume Fraction

Figure 4.15 Specific Energy Absorbed for Sandwich Structures with IPC Foam Core

63
Failure observed in IPC core sandwich structures typically manifested in tensile

failure of the core and lower face sheet. The failure would begin with a crack in the core

in the tensile region. The bottom aluminum face would then yield in the area below the

loading roller. This would lead to a core to face debond in the yielded area, and finally a

crack propagating from the debond area upwards towards the compression zone of the

core. Due to the spontaneous nature of failure initiation, it could not be confirmed in

static tests if a local face-core debonding preceded and contributed to the core crack

initiation or vice versa. However, from observing damaged specimens and from impact

tests (see Section 6.5), it is believed that core crack initiation in the tensile region was

likely the first event leading to failure. Figure 4.16 shows images of a failed IPC foam

core sandwich. Note that the bottom face sheet has undergone yielding underneath the

location of the loading roller. The whole structure remains intact more than those with

other core types because of the metallic ligaments acting as bridging agents. Figure 4.17

shows some magnified images of the crack. In image (a), slight face-core and interphase

separation are observed as well as a slight crack in the face moving outward from the

core interface. In image (b) an aluminum ligament has yielded and failed in the tensile

zone. Image (c) shows some of the aluminum ligaments that have separated from the SF,

but serve as bridging agents, holding the structure together.

(a) (b)

Figure 4.16 (a) Failed IPC Foam Core Sandwich with (b) Yielded Bottom Face

64
(a) (b) (c)

Figure 4.17 Magnified Images of Crack Surfaces Near the Lower Face Sheet

The behavior of IPC core sandwich structure resembles a bilinear response. At

the transition point, the behavior deviates from the initial linear elastic region to the

nonlinear portion. The response of a typical specimen at each volume fraction is

compared to the theoretical mid-span displacement calculation from Section 1.5.1 [16].

The load histories are used to calculate the theoretical displacements, along with

geometric and elastic parameters. To theoretically calculate linear elastic mid-span

displacements, the quasi-static elastic properties used for the IPC material measured by

Jhaver [17] under compression were derived from the stress-strain responses shown in

Figure 4.18. The values are shown in Table 4.5. These comparisons of predicted elastic

behavior and experimental mid-span deflection are shown in Figure 4.19 to 4.21.

65
Figure 4.18 Compressive Response of IPC [17]

Table 4.5 IPC Quasi-Static Elastic Properties [17]

Elastic Modulus
Material
(MPa)
IPC20 2123
IPC30 1852
IPC40 1702

Material: IPC20
6000 Temperature: 70°F
Specimen: FLEX-10
Cast: 30Sept2009
Test: 21Apr2010
Rate: 0.025 in/min
5000

4000
Load (N)

Measured
3000

Elastic
Prediction
2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5
Displacement (mm)

Figure 4.19 Predicted vs. Experimental Mid-Span Deflection, IPC20

66
Material: IPC30
6000 Temperature: 70°F
Specimen: FLEX-16
Cast: 15Apr2010
Test: 09Jun2010
Rate: 0.025 in/min
5000

4000
Load (N)

Measured
3000

Elastic
Prediction
2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Displacement (mm)

Figure 4.20 Predicted vs. Experimental Mid-Span Deflection, IPC30

Material: IPC40
5000 Temperature: 70°F
Specimen: FLEX-14
Cast: 07Oct2009
Test: 21Apr2010
Rate: 0.025 in/min
4000

3000
Load (N)

Measured

2000 Elastic
Prediction

1000

0
0.0 0.5 1.0 1.5 2.0 2.5
Displacement (mm)

Figure 4.21 Predicted vs. Experimental Mid-Span Deflection, IPC40

67
The theoretical and experimental values deviate from each other very early on for

all volume fractions of microballoons in the SF. The predictions exhibit greater stiffness

than measurements. Note that since the theoretical response is only for the linear elastic

case, the response after the transition point cannot be compared. Furthermore, the

deviations in the elastic region are possibly due to the fact that the elastic modulus for

IPC foam core is assumed to be the same in tension as in compression. The elastic

stiffness of IPC foam under tension is likely lower than the compression counterpart due

to relatively weak bonding between aluminum ligaments and the SF foam. Furthermore,

the lower stiffness of the structure from early on in the loading history suggests that SF

and aluminum ligaments debond in the tensile region quite early on, causing the

nonlinear behavior. Yet, the global integrity of the core continues until much later on in

the loading history. Figure 4.22 shows a load-displacement overlay of the three

specimens used as examples, with each containing a different volume fraction (Vf) of

microballoons in the SF. The transition point is virtually the same for all three cases.

After the transition, nonlinear behavior is similar across the three different cases, with the

30% Vf having both the greatest energy and specific energy absorbtion capacity. The

transition point can be considered comparable to a yield point, beyond which

irrecoverable plastic deformation occurs.

68
9000 Material: IPC
Temperature: 70°F
Overlay

8000

7000 20%

6000 IPC30
FLEX-10 (IPC20)

30%
Load (N)
5000
IPC20
FLEX-16 (IPC30)
4000
IPC40
40%

3000
FLEX-14 (IPC40)

2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Crosshead Displacement (mm)

Figure 4.22 Overlay of Measured Load-Displacement of Sandwich Structures with IPC

Foam Core with Different Vf of Microballoons in the SF

4.3 Summary

While the IPC foam core sandwiches experienced slightly lower load levels than

their syntactic foam core counterparts (comparisons are discussed in Chapter 7), the

deflection at failure was greatly improved, thus improving the overall strain energy

absorption. There was also the new addition of plastic nonlinear behavior in the IPC core

sandwiches that was not seen in the SFS sandwiches. Also of note is that these sandwich

beams did not fail abruptly, but rather developed a core crack and face sheet yielding.

The aluminum foam ligaments acted as bridges holding the SF and hence the sandwich

together, even when it could no longer take any significant load. This could be important

from an overall structural aspect, as one would not want a structure to catastrophically

69
fail after, say, being struck by a projectile or flying debris. It would be better to have a

weakened area on an intact structure than to have a catastrophic rupture in that structure.

Considering that part of the reason for IPC core sandwich failure resides in the

IPC itself and that the sandwiches with SF core typically failed due to face-core

debonding, it was considered that if we were somehow able to improve the face-core

bonding while using syntactic foam in the core, a superior material might emerge. Using

syntactic foam in the core rather than IPC would prevent the interphase separation that

acts as microcracks. This interphase separation in the tensile region is illustrated in

Figure 4.23.

SF Alum foam Interphase debond

(a) (b)

Figure 4.23 Illustration of IPC (a) Before and (b) After Interphase Separation between

Syntactic Foam and Aluminum Ligaments

In order to achieve this improved bonding, similar aluminum face sheets were

used with one important difference: each face sheet had ~2mm of brazed aluminum foam

on one side. This narrow (thin) layer of aluminum foam would act as an extremely

roughed surface for bonding, acting like little fingers gripping the core and hopefully

70
preventing a premature face-core shear failure allowing the SF to perform its role of

keeping the two faces apart for effective flexural performance. This new type of face

sheet is referred to as a “graded” face sheet and will be discussed in the next chapter.

71
CHAPTER 5

FLEXURAL CHARACTERISTICS OF SYNTACTIC FOAM CORE

SANDWICH STRUCTURES WITH GRADED FACE SHEETS

5.1 Experimental Setup

This part of the research explored sandwich structures with SF foam core and

graded (or tailored) face sheets. Experimental characterization in bending was carried

out using the same setup described for structures with syntactic foam core and IPC foam

core in Chapter 3 and 4, respectively.

The 0.813mm thick face sheets had a ~2 mm layer of brazed aluminum foam on

one side to act as a “roughened face” that would more readily dig into the SF core to

transfer the load throughout the specimen. This new type of face sheet would also

hopefully prevent the face-core debond that caused premature failure seen in the earlier

syntactic foam core sandwiches. Additionally, there would not be a great amount of

metallic foam, as the failure mechanisms in the IPC foam core stemmed from the

interphase separation between the SF and aluminum ligaments. Figure 5.1 shows an

image of the syntactic foam core sandwich with graded face sheets (identified as SFS-b

hereafter). Note that the regular diagonal marks easily observed on the face sheets are

tool marks.

72
0.813mm
Aluminum foam

SF core

AL6061
~2mm face sheet

Figure 5.1 Syntactic Foam Core Sandwich Structure with Graded Face Sheets (SFS-b)

5.2 Quasi-Static Tests

Flexural tests similar to those described in the previous chapters were carried out

for syntactic foam core sandwiches with graded face sheets and load-deflection responses

were measured. The tests were aimed towards investigating how a tailored or graded

face sheet could prevent premature shear separation between the face and the core. The

load-deflection responses for SFS-b sandwiches are shown in Figures 5.2 to 5.4, and the

results compiled from these tests are summarized in Tables 5.1 to 5.3. The different

materials with graded face sheets and 20%, 30%, and 40% volume fractions (Vf) of

microballoons in the SF core will henceforth be referred to as SFSb20, SFSb30, and

SFSb40, respectively. The “b” is simply a labeling convenience, with no greater

significance. The first type of syntactic foam sandwich face sheet (sanded)/SF core

combination is identified simply as “SFS”, and the second type of face sheet (graded)/SF

core combination is identifed as “SFS-b”.

73
9000 Material: SFSb20
Temperature: 70°F
Overlay

8000

7000

6000
FLEX-57
(SFSb20)
Load (N)
5000
FLEX-58
(SFSb20)
4000

3000 FLEX-61
(SFSb20)

2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Crosshead Displacement (mm)

Figure 5.2 Overlay of Measured Load-Deflection Data for Three Different SF Core

Sandwich Beams with Graded Face Sheets (SFSb20)

9000 Material: SFSb30


Temperature: 70°F
Overlay

8000

7000

6000 FLEX-55
(SFSb30)
Load (N)

5000 FLEX-56
(SFSb30)

4000 FLEX-59
(SFSb30)

3000 FLEX-60
(SFSb30)

2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Crosshead Displacement (mm)

Figure 5.3 Overlay of Measured Load-Deflection Data for Four Different SF Core

Sandwich Beams with Graded Face Sheets (SFSb30)

74
9000 Material: SFSb40
Temperature: 70°F
Overlay

8000

7000

6000 FLEX-63
(SFSb40)
Load (N)

5000 FLEX-64
(SFSb40)

4000 FLEX-65
(SFSb40)

3000 FLEX-66
(SFSb40)

2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Crosshead Displacement (mm)

Figure 5.4 Overlay of Measured Load-Deflection Data for Four Different SF Core

Sandwich Beams with Graded Face Sheets (SFSb40)

Table 5.1 TPB SFSb20 Test Results

Crosshead
Specimen Speed Energy Specific Energy Peak Deflection Transition Transition
Number Type VF(%) (in/min) Absorbed (J) Absorbed (J/kg) Load (N) at Failure (mm) Load (N) Deflection (mm)
57 SFSb 20 0.025 11.59 145.4 8173 2.18 5876 0.95
58 SFSb 20 0.025 11.57 144.7 8156 2.16 5551 0.85
61 SFSb 20 0.025 11.47 143.1 8027 2.18 5960 0.98
NUM OF SAMPLES 3 3 3 3 3 3
AVERAGE 11.54 144.4 8119 2.18 5796 0.93
STANDARD DEVIATION 0.06 1.2 80 0.01 216 0.07

Table 5.2 TPB SFSb30 Test Results

Crosshead
Specimen Speed Energy Specific Energy Peak Deflection Transition Transition
Number Type VF(%) (in/min) Absorbed (J) Absorbed (J/kg) Load (N) at Failure (mm) Load (N) Deflection (mm)
55 SFSb 30 0.025 11.16 152.5 7581 2.21 5885 0.98
56 SFSb 30 0.025 11.85 160.8 7469 2.41 5618 1.03
59 SFSb 30 0.025 12.21 167.6 7570 2.39 5738 0.98
60 SFSb 30 0.025 12.81 175.2 7505 2.49 5658 0.98
NUM OF SAMPLES 4 4 4 4 4 4
AVERAGE 12.01 164.0 7531 2.37 5725 0.99
STANDARD DEVIATION 0.69 9.7 54 0.12 118 0.03

75
Table 5.3 TPB SFSb40 Test Results

Crosshead
Specimen Speed Energy Specific Energy Peak Deflection Transition Transition
Number Type VF(%) (in/min) Absorbed (J) Absorbed (J/kg) Load (N) at Failure (mm) Load (N) Deflection (mm)
63 SFSb 40 0.025 10.57 157.3 6916 2.26 5137 0.89
64 SFSb 40 0.025 11.22 168.4 6328 2.44 4964 0.91
65 SFSb 40 0.025 11.04 166.4 6738 2.36 5222 0.91
66 SFSb 40 0.025 10.76 163.2 6479 2.34 5035 0.91
NUM OF SAMPLES 4 4 4 4 4 4
AVERAGE 10.90 163.8 6615 2.35 5090 0.91
STANDARD DEVIATION 0.29 4.8 263 0.07 113 0.01

In the case of the SFS sandwiches, the load-deflection behavior was essentially

linear elastic in nature (see Chapter 3) with face-core debonding leading to premature

ultimate failure. Moreover there was also substantial scatter in peak load and deflection

at failure. However, the SFS-b sandwiches clearly do not show any face-core debonding,

primarily attributed to the graded/tailored faces penetrating the core over a small spatial

distance adjacent to the face sheet, muting the intensity of shear stresses. This allows the

sandwich structure in general and the SF core in particular to contribute to the overall

integrity. These can be observed by the nonlinear load-deflection responses in Figures

5.2 to 5.4. Similar to the case with IPC core sandwiches, there is a kink or a “knee” in

each plot (the transition point) beyond which noticeable nonlinearity prevails up to

ultimate failure. To find the transition point, a line was fit through the data before

transition and another was fit through the data after transition. The point where these two

fitted lines intersect is identified as the transition point. Both the load and deflection at

the transition point were recorded from the graphs. The aluminum foam core sandwiches

exhibit a similar trend of transition from linear to nonlinear behavior. However, unlike

the unfilled aluminum foam core sandwiches, the SFS-b sandwiches are able to absorb

additional load while continuing to deform.

76
The plots and pictures shown below in Figures 5.5 to 5.8 illustrate the progression

of the test at important points of interest. The first photo is the undeformed specimen

used as a reference image for the particular test. The second photo is at point A (in the

plot in Figure 5.5), the transition point, where the behavior deviates from its initial linear

elastic response. The third photo is at the very end of the test once the sample has

fractured. (Analyzed images of this specimen (FLEX-59, SFSb30) are shown in Section

6.4.3).

Material: SFSb30
9000 Temperature: 70°F
Specimen: FLEX-59
Cast: 25Jan2011
8000 Test: 02Feb2011
Rate: 0.025 in/min

7000

6000
B
Load (N)

5000

4000 Measured
A
3000

2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Displacement (mm)

Figure 5.5 Typical SFS-b Load-Deflection Curve: Point A – Transition Point,

Point B – Failure

77
Figure 5.6 Undeformed, P=0 N

Figure 5.7 Point A: Transition Point, P=5127 N

78
Figure 5.8 Point B: Failure

The failure load is the highest for SFSb20, with the ones for SFSb30 and SFSb40

being successively lower. This is a similar trend to that of the SFS material, which had

decreasing peak load averages with increasing volume fraction. The average load at

failure for SFSb20 is 8,119 N, for SFSb30 is 7,531 N, and for IPC40 is 6,615 N. To

compare with the behavior of the aluminum foam core sandwich, the average load at

failure for the SFS-b sandwich is normalized by the plateau load experienced by the

unfilled aluminum foam core sandwich. The average normalized load at failure for

SFSb20 is 20.76, for SFSb30 is 19.26, and for SFSb40 is 16.92. In Figure 5.9 these

values are shown as a histogram. The plateau load for the aluminum foam core sandwich

is the load at which the stress remains nearly constant while able to withstand a large

increase in strain. It is this load that is typically used as the desired design load. Refer to

Figure 3.5 for a typical stress-strain response of this type. The SFS-b sandwiches do not

experience a similar near constant load, but rather a steady increase (hardening behavior)

79
until failure. This load at failure, when compared to the plateau load, is seventeen to

twenty-one times higher than the design load for unfilled aluminum foam core

sandwiches. The SFS-b sandwiches show superior load bearing capacity compared to

their unfilled aluminum foam counterpart. SFS-b sandwiches also outperform both SFS

and IPC sandwiches, which will be further discussed in Chapter 7.

25

SFSb

20
Peak Load/ALS Plateau Load

15

10

0
20% 30% 40%
Volume Fraction

Figure 5.9 Normalized Measured Peak Load for Sandwich Structures with Syntactic

Foam Core and Graded Face Sheets

The average value for the mid-span deflection at failure for SFSb20 is 2.18 mm,

for SFSb30 is 2.37 mm, and for SFSb40 is 2.35 mm. To compare with the behavior of

the aluminum foam core sandwich, the average deflection at failure for the SFS-b

sandwich is normalized by the deflection at which the aluminum foam core sandwich

behavior deviates from linearity. The average normalized deflection at failure for

80
SFSb20 is 11.23, for SFSb30 is 12.26, and for SFSb40 is 12.13. In Figure 5.10 these

values are shown as a histogram. Note that within the experimental scatter, there is no

improvement between the 30% and 40% volume fractions although the latter is lighter

overall.

14.0

SFSb
12.0

10.0
δFailure/δALS, nonlinear

8.0

6.0

4.0

2.0

0.0
20% 30% 40%
Volume Fraction

Figure 5.10 Normalized Measured Mid-Span (Load-Point) Deflection at Failure for

Sandwich Structures with Syntactic Foam Core and Graded Face Sheets

The average value for the transition load for SFSb20 is 5,796 N, for SFSb30 is

5,725 N, and for SFSb40 is 5,090 N. As with the peak load, to compare to the behavior

of the aluminum foam core sandwich, the average transition load for the SFS-b sandwich

is normalized by the plateau load experienced by the aluminum foam core sandwich. The

average normalized transition load for SFSb20 is 14.82, for SFSb30 is 14.64, and for

SFSb40 is 13.02. In Figure 5.11 these values are shown as a histogram. This is the best

81
comparison that can be made between syntactic foam and unfilled aluminum foam core

sandwiches, as both have a transition point. With the SFS-b sandwich having a transition

load of around thirteen to fifteen times greater than unfilled aluminum foam sandwiches

and a similar transition deflection, the point at which plasticity begins to set in for the

SFS-b sandwiches is greatly improved. Evidently, there is no distinguishable difference

in the transition load values within the experimental scatter between SFSb20 and

SFSb30, with SFSb40 specimens showing a slightly lower value relative to the two

others. This suggests that face sheet yielding is likely responsible for the transition point

in SFS-b sandwiches as the metallic ligaments of the graded face sheet help prevent shear

failure of the face-core interface.

18
SFSb
16
Transition Load/ALS Plateau Load

14

12

10

0
20% 30% 40%
Volume Fraction

Figure 5.11 Normalized Measured Load at Transition Point for Sandwich Structures

with Syntactic Foam Core and Graded Face Sheets

82
The average value for the transition deflection for SFSb20 is 0.93 mm, for

SFSb30 is 0.99 mm, and for SFSb40 is 0.91 mm. For a comparison, the average

transition deflection for the SFS-b sandwich is normalized by the deflection at which the

aluminum foam core sandwich behavior deviates from linearity. The average normalized

transition deflection for SFSb20 is 4.79, for SFSb30 is 5.11, and for SFSb40 is 4.69. In

Figure 5.12 these values are shown as a histogram. With a four and a half to five times

improvement over unfilled aluminum foam core sandwiches, SFS-b sandwiches are able

to remain in the linear elastic regime much longer before the onset of plasticity.

6.0
SFSb

5.0

4.0
δTransition/δALS, nonlinear

3.0

2.0

1.0

0.0
20% 30% 40%
Volume Fraction

Figure 5.12 Normalized Measured Mid-Span (Load-Point) Deflection at Transition

Point for Sandwich Structures with Syntactic Foam Core and Graded Face Sheets

As before, the strain energy (U) absorbed by a specimen was calculated using the

area under the load-deflection curve from the measured data. That is,

83
δ
U = ∫ P (δ ) d δ .
0

The average values for the strain energy absorbed are plotted in Figure 5.13 as a

histogram. The average for SFSb20 is 11.54 J, for SFSb30 is 12.01 J, and for SFSb40 is

10.90 J.

14.0

SFSb
12.0

10.0
Energy Absorbed (J)

8.0

6.0

4.0

2.0

0.0
20% 30% 40%
Volume Fraction

Figure 5.13 Energy Absorbed for Sandwich Structures with Syntactic Foam Core and

Graded Face Sheets

In order to compare among varying volume fractions of microballoons in the SF,

specific energy (J/kg) was calculated for each specimen. The strain energy absorbed was

divided by the mass of the particular specimen. The mass was determined by,

84
mtot = ρcoreVcore + ρ facesV faces .

The details of material densities used are in Table 4.5. The results for specific

strain energy are shown in Figure 5.14. The average for SFSb20 is 144.4 J/kg, for

SFSb30 is 164.0 J/kg, and for SFSb40 is 163.8 J/kg. Comparing the specific strain

energies of SFS-b and IPC sandwiches (which accounts for the weight difference in

removing the aluminum foam in the core), SFSb20 had a 71% improvement over its IPC

counterpart, SFSb30 had a 76% improvement, and SFSb40 had a 118% improvement.

With syntactic foam core outperforming IPC foam core for all volume fractions as well as

being lighter with the removal of the aluminum foam matrix, it can be concluded that

SFS-b sandwiches are far superior to a similar configuration with IPC foam core in their

energy absorbing capabilities. It is now apparent that the best core type is syntactic foam

rather than IPC. However, the face sheets simply need to be tailored to penetrate the SF

core in the face-core region to reap the full benefits of the face and of the core by

preventing premature face-core shear failure yet allowing effective load transfer between

the face and the core.

85
200
SFSb
180

160

Specific Energy Absorbed (J/kg)


140

120

100

80

60

40

20

0
20% 30% 40%
Volume Fraction

Figure 5.14 Specific Energy Absorbed for Sandwich Structures with Syntactic Foam

Core and Graded Face Sheets

Failure observed in SFS-b sandwiches typically manifested as tensile failure of

the lower face sheet, likely due to a rupture in the SF core, with a crack propagating from

the tensile zone upwards towards the compression zone. Due to the dynamic nature of

failure initiation, it could not be confirmed if lower face sheet yielding preceded and

contributed to the core crack initiation or vice versa. Images of a failed SFS-b sandwich

are shown in Figure 5.15. Also, some magnified images of a failed SFS-b sandwich are

shown in Figure 5.16. In Figure 5.16 notice in (a) and (b) that the face sheets have

appeared to again fail in tension after yielding. Also notice the abrupt breaks in the

metallic ligaments in (c), as well as the SF still adhered to the metallic ligaments,

indicating good adhesion due to silane treatment. Though no images captured the actual

failure event, the abrupt face sheet separation seen in Figure 5.16 (a) and (b) indicates

86
that the failure initiated in the core, expanding downward to the face sheet and upward to

the compression zone.

(a)

(b)

(c)

Figure 5.15 Failed SFS-b Specimen with (a) Side, (b) Top Face, and (c) Bottom Face

Views

87
(a) (b) (c)

Figure 5.16 Magnified Images (a), (b) of Failure Surfaces and (c) From Below

The behavior of SFS-b sandwiches resembles a bilinear response. At the

transition point, the behavior deviates from the initial linear elastic regime to the

nonlinear regime. The response of a typical specimen at each volume fraction is

compared to the theoretical mid-span deflection calculation from Section 1.5.1 [16]. The

load histories are used to calculate the theoretical deflections, along with geometric and

elastic parameters. To calculate predicted deflections based on linear elastic material

behavior, the quasi-static elastic properties used for the syntactic foam found by Jhaver

[17] were used once again. (Refer back to Figure 3.7 for the stress-strain responses from

which they were derived and Table 3.2 for the properties.) There is good agreement

between the theoretical and experimental values in the initial linear elastic regime for all

volume fractions. The predictions exhibit similar stiffness compared to those seen in the

measured data. Note that since the theoretical response is only for the linear elastic case,

the response after the transition point cannot be compared. The deviations after the

elastic region are possibly due to the fact that the elastic modulus for syntactic foam core

88
is assumed to be the same in tension as in compression. These comparisons between the

predicted and experimental behavior are shown below in Figure 5.17 to 5.19.

Material: SFSb20
9000 Temperature: 70°F
Specimen: FLEX-57
Cast: 23Jan2011
8000 Test: 10Feb2011
Rate: 0.025 in/min

7000

6000
Load (N)

5000
Measured

4000
Elastic
Prediction
3000

2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5
Displacement (mm)

Figure 5.17 Predicted vs. Experimental Mid-Span Deflection, SFSb20

Material: SFSb30
9000 Temperature: 70°F
Specimen: FLEX-59
Cast: 25Jan2011
8000 Test: 02Feb2011
Rate: 0.025 in/min

7000

6000
Load (N)

5000
Measured

4000
Elastic
Prediction
3000

2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Displacement (mm)

Figure 5.18 Predicted vs. Experimental Mid-Span Deflection, SFSb30

89
Material: SFSb40
8000 Temperature: 70°F
Specimen: FLEX-65
Cast: 01Feb2011
7000 Test: 10Feb2011
Rate: 0.025 in/min

6000

5000

Load (N)
Measured
4000

Elastic
3000 Prediction

2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5
Displacement (mm)

Figure 5.19 Predicted vs. Experimental Mid-Span Deflection, SFSb40

Figure 5.20 shows a load-displacement overlay of the three specimens used as

examples, with each containing a different volume fraction of microballoons in the SF

core.

9000 Material: SFS-b


Temperature: 70°F
Overlay

8000 SFSb20

SFSb30
7000 20%

6000 SFSb40
FLEX-57 (SFS20)

30%
Load (N)

5000
FLEX-59 (SFS30)
4000
40%

3000
FLEX-65 (SFS40)

2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Crosshead Displacement (mm)

Figure 5.20 Overlay of Measured Load-Displacement of SF Core Sandwich Structures

with Graded Face Sheets and Different Vf of Microballoons in the SF

90
The transition deflection is virtually the same for all the three cases. After the

transition, nonlinear behavior is similar across the different cases with decreasing

stiffness as volume fraction increases. While SFSb30 has the greatest energy absorbing

capacity, the 30% and 40% Vf cases have the same specific energy absorbing capacity.

So if energy absorption is the driving factor, perhaps the 30% Vf case might be the best

core material. However, if both weight saving and energy absorbing capacity are both

important attributes, then the 30% and 40% Vf would both warrant consideration.

5.3 Summary

With the addition of the graded face sheets, the syntactic foam core sandwiches

(SFS-b) outperformed IPC foam core sandwiches in all metrics (comparisons are

discussed in Chapter 7). Not only did it outperform the IPC core sandwiches, but without

the metallic matrix in the core, the syntactic foam core sandwiches are lighter. It is easy

to see now that the IPC core sandwiches, though substantially better than conventional SF

core sandwiches, was an inferior material for flexural loading conditions due to its

weakness under the tensile conditions. As a single continuous phase the syntactic foam

does not experience the interphase separation that IPC foam does that essentially acts as

micro-defects, ultimately leading to tensile failure. The core integrity is better with the

syntactic foam. These results were not seen with the first type of SFS sandwiches. The

poor bond between the sanded face sheets and the core was not able to effectively

transfer load between the face and the core. Failure would occur prematurely when the

91
face sheet and core would debond, leaving the core to bear the entire load. The graded

face sheets in the SFS-b sandwiches remedied this problem, allowing for effective load

transfer throughout the specimen while minimizing face-core shear jump.

92
CHAPTER 6

FULL-FIELD DEFORMATION MEASUREMENTS USING DIGITAL IMAGE

CORRELATION METHOD

6.1 Digital Image Correlation (DIC) Principle

Digital image correlation (DIC) is a relatively new optical method and is based on

the principle of locating a point (or region of interest) in the deformed image of an object

relative to its location in the undeformed state [20,21,22,23]. The object surface is

decorated with a random pattern (in this case a random black/white speckle). The

decoration is illuminated using white light sources to produce diffusely reflected light

from the specimen surface. The light intensity (grayscales) of the surface is used for the

pattern matching process. The recording of the surface patterns is done using a digital

camera with a relatively high pixel count (typically on the order of one to ten megapixels)

to achieve good spatial resolution.

During a typical digital imaging process, an analog intensity field I(x,y) is

converted into a discrete field D(x,y). The output from an image is typically a two

dimensional array of intensity levels (D(i,j) consisting of i rows and j columns). If

discrete intensity fields D and D’ are available for two load levels (or, reference and

deformed states), then a subset d from D (say, a 15 x 15 pixel subimage from a 1000 x

93
1000 pixel image) is chosen and its location in D’ (also a 1000 x 1000 pixel image) is

searched.

A typical random speckle pattern in undeformed and deformed states is shown in

Figure 6.1. The highlighted box on the undeformed image indicates the area to track

during correlation. The box on the deformed image indicates the same area after

undergoing deformation. A schematic of the planar deformation is depicted in Figure

6.2.

(a) (b)

Figure 6.1 (a) Undeformed and (b) Deformed Speckle Images

94
UNDEFORMED SUBIMAGE

P
∆x • u
O
v

• P’
O’
∆y
DEFORMED SUBIMAGE

Figure 6.2 Tracked Subimage for Image Correlation

A subimage with O as the center is displaced to O’ after deformation. If u and v

denote displacement components in the x- and y- directions, the process of deformation

can be expressed using affine coordinate transformations to relate O’(x’,y’) with O(x,y).

This requires an accurate interpolation method to express intensity variation within the

subimage.

For a subimage centered at point O in the undeformed state, the discretely

sampled intensity and continuously interpolated intensity at O and a neighboring point P

at positions (x,y) and (x+dx,y+dy) can be written as

D (O ) = D ( x, y )
D ( P ) = D ( x + dx, y + dy )

The differential distances between neighboring points O and P are represented by

dx and dy. After deformation the intensity patterns recorded are expressed using

x ' = x + u ( x, y ); y ' = y + v ( x, y ) .

95
So,

D '( x ', y ') = D ( x + u ( x, y ), y + v( x, y ))


D '( x '+ dx ', y '+ dy ') = D[ x + dx + u ( x + dx, y + dy ), y + dy + v( x + dx, y + dy )]
∂u ∂u ∂v ∂v
D '( x '+ dx ', y '+ dy ') = D[ x + u ( x, y ) + (1 + ) dx + dy, y + v( x, y ) + dx + (1 + ) dy ]
∂x ∂y ∂x ∂y

If the subimage is sufficiently small and displacement gradients are therefore

∂u ∂u ∂v ∂v
nearly constant within the subimage, one can obtain u, v, , , , for each
∂x ∂y ∂x ∂y

subimage. This is generally done using one of the following measures: (1) magnitude of

intensity difference, (2) sum of the squares of intensity value difference, (3) normalized

cross-correlation, or (4) cross-correlation. The most commonly used is normalized cross-

correlation where the following quantity,

∑ D '(O ') D(O)


C NCC = i
,
∑ D '(O ')2
i
∑ D(O)2
i

is maximized to provide the optimal estimation of all the six displacements and

displacement gradients. A number of optimization methods can be used for this purpose

and include the so-called Newton-Raphson, coarse-fine, and Levenburgh-Marquardt

methods. Levenburgh-Marquardt method has been demonstrated to have superior

convergence characteristics and is as fast as the Newton-Raphson method.

96
6.2 ARAMIS™ Image Analysis Software

Full field deformation and strain analysis was performed using the 2D image

correlation technique utilizing a spray-on speckle pattern. The software used for the

image post-processing was ARAMIS™ (GOM mbH, Germany). It is able to provide full

field measurements using a photogrammetric method [24] as the specimen undergoes

deformation during a test. A Nikon D100 digital camera was used during experiments to

record the grayscales (at 10-bit resolution) corresponding to the specimen deformation.

Although the camera had a 6 megapixel resolution, during time-lapse photography only

1500 x 1000 pixels were used for easy throughput of grayscale information into a laptop

using Nikon Capture software (Version 4.0). The entire image is regularly divided into a

number of smaller sub-images (facets). The photogrammetric principles used to evaluate

the images track facets in each successive image. From each facet, a single measurement

results after computation. The facet size used (15 x 15 pixels) was a good compromise

between accuracy and computational time. The goal is to have a good representation of

the speckle pattern within the facet. This digital image processing provides full field

displacement and strain contours throughout the test. Furthermore, not only can this be

implemented in quasi-static situations, but also in dynamic situations [25].

∂u ∂u ∂v ∂v
The deformation gradient tensor F (comprised of , , , ) creates a
∂x ∂y ∂x ∂y

functional connection of the coordinates of the deformed points Pv,i with the coordinates

of the undeformed points Pu,i [26]. They are connected through the relation

Pv ,i = u j + F gPu ,i ,

97
where,

Pv ,i = Deformed point coordinates

Pu ,i = Undeformed point coordinates

uj = Rigid body translation

The different coordinate systems defined through the calculation process are

shown in Figure 6.3. The coordinates of the points (e.g. pu and pv in the figure) are

calculated in the global x-y coordinate system. For 2D analysis, the local undeformed

coordinate system x’-y’ is parallel to x-y, but moved to coincide with the point pu. The

x”-y” coordinate system for the strain calculation is independent of rigid body movement

and rotation.

Figure 6.3 Coordinate System Definition

To calculate the deformation gradient tensor F, the coordinates of each point must

be known both in the undeformed and deformed state. This tensor can be interpreted as

98
an affine transformation which transforms a unit square into a generic quadrilateral. In

order to calculate F for a point, a number of neighboring points are needed. For this

calculation, a homogeneous state of strain is assumed for the set of adjacent points. The

width of the field is the reference length for evaluating strain. (For example in Figure 6.4,

the width of the field is three points or pixels. The reference length is then three pixels.)

The number of surrounding points can be changed, but a 3x3 neighborhood is shown in

Figure 6.4 below.

Figure 6.4 3x3 Point Neighborhood for 2D Strain Calculation

6.3 Experimental Setup

Section 3.1 details the TPB test setup used in conjunction with DIC method and

time-lapse photography. The impact test counterparts were carried out using the speckle

pattern and camera and are detailed in Section 6.5.

99
6.4 Optical Measurement Results

Using built-in algorithms in ARAMIS™, displacement and strain fields were

computed and illustrated. With this capability, it is relatively simple to obtain local

displacement and strain information, rather that relying on the global load and

displacement measurements made using the testing machine. To ensure the ARAMIS™

results are reliable, local and global displacement measurements were compared for each

specimen just below the load roller (the machine displacement data given is for the

crosshead). An example of this is shown in Figure 6.5. This is one example, with every

specimen compared and shown in Appendix A. Since the camera recorded images in ~2

second intervals, there is understandably a small lag in the ARAMIS™ displacements.

Material: IPC30
6000 Temperature: 70°F
Specimen: FLEX-16
Cast: 15Apr2010
Test: 09Jun2010
Rate: 0.025 in/min
5000

4000
Load (N)

3000 Measured

ARAMIS

2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Displacement (mm)

Figure 6.5 Global/Local Displacement Comparison, FLEX-16

100
Also, to check the strain calculations, a sample (129mm length x 5.7mm width x

21mm height) of PMMA (E = 3 GPa) was tested in three-point bending and the strains

calculated using ARAMIS™. These dimensions were chosen to match the planar

dimensions of the sandwich specimens under investigation. The comparison is shown in

Figure 6.6. The measured load was used to calculate the maximum tensile stress using

My σx
σx = and theoretical elastic strain was calculated using ε x = . The optical
I x = 0; y = c E

measurement of strain and the theoretical prediction are plotted in Figure 6.6, with good

agreement between the two being evident. The error bars are determined from the plot in

Figure 6.7. Also plotted are strain (εx) and shear strain (εxy) variations, in Figures 6.7 and

6.8, respectively, along the beam height (y) at x=25mm (x=0 is the mid-span of the

VQ τ
beam). Theoretical shear strain is calculated using τ = and ε xy = . A rather good
Ib 2G

agreement is seen in the εx plot. Note the strange nonlinear behavior at the top and

bottom of the sample. Considering the known behavior of PMMA, it can be concluded

that any optical measurements around the top and bottom of the specimen have an

increased amount of error. This plot shows a maximum error of approximately 600

µstrains, which was determined to be the error associated with optical strain

measurements. The εxy plot trend was the same as the elastic prediction with an error of

approximately 1000 µstrains.

101
Material: PMMA
70 Temperature: 70°F
Specimen: PMMA-1
Test: 04Mar2011
Rate: 0.025 in/min
60

50

Stress (MPa)
40
Measured

30 ARAMIS

20

10

0
0.0 0.5 1.0 1.5 2.0 2.5
Strain (%)

Figure 6.6 Strain Calibration Using PMMA

Material: PMMA
Temperature: 70°F
Rate: 0.025 in/min
9.2 Load: 1,650 N

6.7

4.2

1.7
y (mm)

Elastic
Predicition
-0.8
ARAMIS

-3.3

-5.8

-8.3

-10.8
-1.50 -1.00 -0.50 0.00 0.50 1.00 1.50
ε x (%)

Figure 6.7 PMMA Strain (εx) Along the Height of the Beam

102
Material: PMMA
Temperature: 70°F
Rate: 0.025 in/min
9.2 Load: 1,650 N

6.7

4.2

1.7
y (mm)

Elastic
Prediction
-0.8
ARAMIS

-3.3

-5.8

-8.3

-10.8
0.000 0.001 0.002 0.003 0.004 0.005 0.006 0.007
ε xy (rad)

Figure 6.8 PMMA Shear Strain (εxy) Along the Height of the Beam

6.4.1 Syntactic Foam (SFS) Results

The series of images shown in Figures 6.9 to 6.17 are typical ARAMIS™ results

for tests conducted on SFS sandwiches. Each of these experiments are summarized in

sets of three images: the first image corresponding to a point in the linear response

regime, the second image corresponding to just before failure, and the third image

corresponding to just after failure. Note the presence of strain concentration due to the

loading roller in particular. Though it is hard to conclusively determine given the low

amount of deflection in the beam, the u-displacement fields are similar to what is

expected from the beam theory. Note the units: u-displacement is measured in

millimeters, εx is measured in percent, and εxy is measured in strains. Also note that the

correlated images contain twenty contours. The u-displacement increment is 50µm, εx

increment is 0.25%, and εxy increment is 4,000 µstrains.

103
y

x
P=2346N

P=4946N

P=2869N

Figure 6.9 ARAMIS™ Results for SFS20 (FLEX-38, u)

104
y

x
P=2346N

P=4946N

P=2869N

Figure 6.10 ARAMIS™ Results for SFS20 (FLEX-38, εx)

105
y

x P=2346N

P=4946N

P=2869N

Figure 6.11 ARAMIS™ Results for SFS20 (FLEX-38, εxy)

106
y

x P=2236N

P=5468N

P=3618N

Figure 6.12 ARAMIS™ Results for SFS30 (FLEX-39, u)

107
y

x P=2236N

P=5468N

P=3618N

Figure 6.13 ARAMIS™ Results for SFS30 (FLEX-39, εx)

108
y

x P=2236N

P=5468N

P=3618N

Figure 6.14 ARAMIS™ Results for SFS30 (FLEX-39, εxy)

109
y

x P=2069N

P=4080N

P=2459N

Figure 6.15 ARAMIS™ Results for SFS40 (FLEX-24, u)

110
y

x P=2069N

P=4080N

P=2459N

Figure 6.16 ARAMIS™ Results for SFS40 (FLEX-24, εx)

111
y

x P=2069N

P=4080N

P=2459N

Figure 6.17 ARAMIS™ Results for SFS40 (FLEX-24, εxy)

112
6.4.2 IPC Results

The series of images shown in Figures 6.18 to 6.26 are typical optical results for

tests conducted on IPC sandwiches. In each of these sets of four images, the first image

corresponds to a point in the linear portion of the loading regime, the second image

corresponds to a point in the nonlinear regime, the third image corresponds to the point

just before failure, and the fourth image corresponds to the point just after failure. Notice

that with the IPC, the strain concentration effects due to the rollers are relatively muted,

as the interpenetrating core does a good job of evenly distributing the load. Again, u-

displacement fields qualitatively resemble the ones expected from the beam theory.

113
y

x P=824N

P=3536N

P=4922N

P=3905N

Figure 6.18 ARAMIS™ Results for IPC20 (FLEX-10, u)

114
y

x P=824N

P=3536N

P=4922N

P=3905N

Figure 6.19 ARAMIS™ Results for IPC20 (FLEX-10, εx)

115
y
P=824N
x

P=3563N

P=4922N

P=3905N

Figure 6.20 ARAMIS™ Results for IPC20 (FLEX-10, εxy)

116
y

x P=1097N

P=3997N

P=5680N

P=3741N

Figure 6.21 ARAMIS™ Results for IPC30 (FLEX-16, u)

117
y

x
P=1097N

P=3997N

P=5680N

P=3741N

Figure 6.22 ARAMIS™ Results for IPC30 (FLEX-16, εx)

118
y

x P=1097N

P=3997N

P=5680N

P=3741N

Figure 6.23 ARAMIS™ Results for IPC30 (FLEX-16, εxy)

119
y

x P=950N

P=2845N

P=4442N

P=3159N

Figure 6.24 ARAMIS™ Results for IPC40 (FLEX-14, u)

120
y

x P=950N

P=2845N

P=4442N

P=3159N

Figure 6.25 ARAMIS™ Results for IPC40 (FLEX-14, εx)

121
y

x P=950N

P=2845N

P=4442N

P=3159N

Figure 6.26 ARAMIS™ Results for IPC40 (FLEX-14, εxy)

122
6.4.3 Syntactic Foam (SFS-b) Results

The series of images shown in Figures 6.27 to 6.35 are typical ARAMIS™ results

for tests conducted on syntactic foam core sandwiches with graded face sheets (SFS-b).

In each of these sets of three images, the first image corresponds to a point in the linear

regime, the second image corresponds to a point in the nonlinear regime, and the third

image corresponds to the point just before failure. In the SFS-b sandwiches, the

concentration effects due to the rollers once again dominate. Also of note is that the

shear strains are higher compared to that of the SFS sandwiches due to the more efficient

load transfer throughout the structure enabled by the graded face sheets. Once again, the

u-displacement fields are what you would expect from the beam theory.

123
y

x P=3491N

P=7187N

P=8155N

Figure 6.27 ARAMIS™ Results for SFSb20 (FLEX-57, u)

124
y

x P=3491N

P=7187N

P=8155N

Figure 6.28 ARAMIS™ Results for SFSb20 (FLEX-57, εx)

125
y
P=3491N
x

P=7187N

P=8155N

Figure 6.29 ARAMIS™ Results for SFSb20 (FLEX-57, εxy)

126
y

x P=2990N

P=6717N

P=7559N

Figure 6.30 ARAMIS™ Results for SFSb30 (FLEX-59, u)

127
y

x P=2990N

P=6717N

P=7559N

Figure 6.31 ARAMIS™ Results for SFSb30 (FLEX-59, εx)

128
y

x P=2990N

P=6717N

P=7559N

Figure 6.32 ARAMIS™ Results for SFSb30 (FLEX-59, εxy)

129
y

x P=2207N

P=6054N

P=6735N

Figure 6.33 ARAMIS™ Results for SFSb40 (FLEX-65, u)

130
y

x P=2207N

P=6054N

P=6735N

Figure 6.34 ARAMIS™ Results for SFSb40 (FLEX-65, εx)

131
y

x P=2207N

P=6054N

P=6735N

Figure 6.35 ARAMIS™ Results for SFSb40 (FLEX-65, εxy)

132
6.5 Impact Tests

The impact tests conducted in this work were only conducted on IPC and SFS-b

types using a single volume fraction (30%) of microballoons in the SF. The purpose of

these tests was to further elaborate on the failure mechanisms in the quasi-static case, as

well as begin to understand dynamic characteristics of these sandwich architectures.

Also, the feasibility of using DIC method in conjunction with a high-speed digital camera

to study impact response of sandwich structures is illustrated.

Impact tests were conducted using an Instron Dynatup 9250-HV drop tower to

deliver low velocity impact to IPC foam core and SFS-b sandwich structures with 30%

volume fraction of microballoons in the SF. The nominal specimen size used for

dynamic testing was 127mm x 25mm x 18.5mm (5” length x 1” width x 0.73” height).

Each face sheet had a nominal thickness of 0.813mm, giving a core thickness of

~16.9mm. The specimen was supported over a 4.5 inch span and was impacted by a

hemispherical impactor of radius 0.5 inch.

Two different cameras were used to capture the dynamic event. The first was a

Phantom v710 digital high-speed camera with a single 1280 x 800 CMOS sensor. This

camera recorded a video of the event and was used to determine the overall impact

response on a large time scale. (Using a resolution of 800 x 600 pixels and a framing rate

of 13,000 frames per second, the total recording duration was ~0.75 seconds.) Images

from the Phantom camera were used to determine the time at which crack initiation was

likely to occur. (To view a video of the impact event created using the Phantom camera,

refer to the disk in the back cover of this thesis.) Using this estimated time, parameters

133
were selected for the second camera, a Cordin Model 550 high-speed digital camera,

which captured the speckle patterns during the dynamic event over a time period of ~600

microseconds. (The Cordin camera is capable of simultaneously recording at much

higher spatial resolutions (1000 x 1000 pixels) and speeds and hence was preferred to the

Phantom camera.) The drop tower has an instrumented tup and a pair of anvils for

recording impact force and reaction force histories, respectively. Two high intensity

flash lamps, triggered by the camera, were used to illuminate the specimen surface. Two

separate computers were also used in the setup. One was used to record the force

histories, and the other was used to control the camera and store the digital images. A

schematic of the test setup with the Cordin camera is shown in Figure 6.36. The setup

consists of (1) high-speed digital camera, (2) drop tower impactor tup, (3) delay

generator, (4) specimen, (5) light sources, (6) DAQ – drop tower, (7) DAQ – camera, (8)

lamp control unit, and (9) drop tower controller.

Region of Interest
(25mm x 25mm)

Figure 6.36 Schematic of Impact Test Setup: (1) high-speed digital camera, (2) drop

tower impactor tup, (3) delay generator, (4) specimen, (5) light sources, (6) DAQ – drop

tower, (7) DAQ – camera, (8) lamp control unit, and (9) drop tower controller

134
A picture of the impact test setup is shown in Figure 6.37 which shows (1) high-

speed camera, (2) drop tower impact tup, (3) specimen, and (4) light source.

4
1

Figure 6.37 Impact Test Setup: (1) high-speed camera, (2) drop tower impactor tup,

(3) specimen, and (4) light source

The high-speed camera employs CCD technology with a high-speed rotating

mirror optical system. The camera is capable of capturing images at a rate of 2,000,000

frames per second at a resolution of 1000 x 1000 pixels for each image. Thirty-two (32)

independent image sensors are positioned circumferentially around a five-facet high-

speed rotating mirror that reflects and sweeps light over the sensors. The 32 sensors

function as individual cameras that operate sequentially by means of internal electronic

triggering. Minor misalignments occur, such as differences in focus and rotational

discrepancies between successive images. This is resolved by grouping the images into

pairs that consist of an undeformed and a deformed image recorded by the same sensor.

135
Before impact a set of 32 images was recorded in the area of interest with the desired

frame rate (50,000 frames per second for this work). After aligning the optics, the

rotating mirror was brought to the desired speed. The camera and lamps were triggered

by the operator. This set of 32 images (one per sensor) of the specimen surface was then

stored for use as undeformed, or reference, images. Maintaining all the same camera

settings, a second set of 32 images was recorded as the specimen experienced the impact

event. During the test, the camera and flash lamps were triggered by the tup as it

contacted a copper strip adhered to the top of the specimen. The first image was

identified, and then sequential pairs were made in the undeformed and deformed sets

(i.e., in the first set of 32 undeformed images, each image has a corresponding deformed

image in the second set). Each of the 32 pairs was analyzed separately. (To view a video

of the impact event created using the Cordin camera, refer to the disk in the back cover of

this thesis.) The impactor velocity used in this work was 5 m/s. The camera viewing

area is shown in Figure 6.38 and was approximately 25mm x 25mm.

Figure 6.38 Camera Viewing Area

IPC foam core sandwiches with a 30% volume fraction of microballoons in the

SF core were impact tested. Figures 6.39 and 6.40 show a sample of correlated images

136
from the Cordin camera on an IPC30 specimen. Shown here are the u-displacement and

normal strain (εx) fields. Note that the correlated images contain twenty contours. The u-

displacement increment is 25µm and εx increments are 0.2%.

t=510µs

t=530µs

t=550µs
Crack
initiation

Figure 6.39 ARAMIS™ Results for IPC30 (DYN-29, u)

137
t=510µs

t=530µs

t=550µs

Crack
initiation

Figure 6.40 ARAMIS™ Results for IPC30 (DYN-29, εx)

The normal (dominant) strain, εx, image preceding visible crack initiation (t=530

µs) contains an anomaly not seen in the previous images in the area where the crack

forms. Also, no significant strain concentration at the impact edge is seen since the

impactor head is hemispherical and the front face of the specimen is far removed from

the specimen mid-plane. Since the crack first forms in the core, then propagates to the

lower face and upper compression zone, it can be concluded that the mechanism likely

138
responsible for both quasi-static and dynamic failure in the IPC foam core sandwiches is

tensile failure of the core.

SFS-b sandwiches with a 30% volume fraction of microballoons in the SF were

also impact tested. Figures 6.41 and 6.42 show a sample of correlated images from the

Cordin camera on an SFSb30 specimen. Shown here are the u-displacement and normal

strain (εx) fields.

t=510µs

t=530µs

t=550µs

Crack
initiation

Figure 6.41 ARAMIS™ Results for SFSb30 (DYN-35, u)

139
t=510µs

t=530µs

t=550µs

Crack
initiation

Figure 6.42 ARAMIS™ Results for SFSb30 (DYN-35, εx)

Similar to the IPC foam core sandwiches, the crack first forms in the tensile

region of the SFS-b sandwich core, only then propagating to the lower face and upper

compression zone. It can be concluded that the initial mechanism likely responsible for

both quasi-static and dynamic failure in the SFS-b foam core sandwiches is also tensile

failure of the core.

Figures 6.43 and 6.44 show plots of the strain energy absorbed by the specimen as

a function of time for the IPC30 and SFSb30 sandwiches, respectively. The results are

140
summarized in Tables 6.1 and 6.2. The Impulse software (for the drop tower)

automatically performs energy calculations using the following relationship,

E (t ) = K (t ) + V (t ) + Ea (t ) = Const

where E(t) is the total energy in the system (which remains constant), K(t) is the kinetic

energy of the drop weight, V(t) is the potential energy of the drop weight, and Ea(t) is the

strain energy absorbed by the specimen up to time t. Graphically, Ea(t) is simply the area

under the load-deflection curve. It is calculated mathematically using the mass (m),

velocity (v(t)), and position (x(t)) of the drop weight (trapezoid integration used to find

v(t) and x(t)) in the equation,

1
Ea (t ) = m vimpact
2
− v 2 (t )  + mgx(t ) .
2

Material: IPC30
16 Temperature: 70°F
Core is Core is Overlay
intact cracked;
14
specimen
continues
12 to absorb
energy
Energy Absorbed (J)

10 Core crack
initiation
DYN-18
8 (IPC30)
DYN-20
(IPC30)

6 DYN-29
(IPC30)

0
0.0 0.5 1.0 1.5 2.0
Time (ms)

Figure 6.43 Strain Energy History, IPC30

141
Material: SFSb30
16 Temperature: 70°F
Overlay

14
Core crack
12
Energy Absorbed (J) initiation

10

DYN-34
8 (SFSb30)
DYN-35
(SFSb30)

6 DYN-36
(SFSb30)

0
0.0 0.5 1.0 1.5 2.0
Time (ms)

Figure 6.44 Strain Energy History, SFSb30

Table 6.1 IPC30 Impact Test Results

Specimen Impact Energy Specific Energy Energy Absorption


Number Type VF(%) Speed (m/s) Absorbed (J) Absorbed (J/kg) Rate (J/ms)
18 IPC 30 5.0 10.8 157.8 12
20 IPC 30 5.0 12.1 178.0 12
29 IPC 30 5.0 11.0 161.1 12
NUM OF SAMPLES 3 3 3
AVERAGE 11.3 165.6 12
STANDARD DEVIATION 0.7 10.8 0

Table 6.2 SFSb30 Impact Test Results

Specimen Impact Energy Specific Energy Energy Absorption


Number Type VF(%) Speed (m/s) Absorbed (J) Absorbed (J/kg) Rate (J/ms)
34 SFSb 30 5.0 8.4 136.5 20
35 SFSb 30 5.0 8.8 143.1 20
36 SFSb 30 5.0 9.2 146.7 20
NUM OF SAMPLES 3 3 3
AVERAGE 8.8 142.1 20
STANDARD DEVIATION 0.4 5.2 0

142
Two major aspects of note are the rate of energy absorption and the total strain

energy absorbed by the specimen. While the SFSb30 sandwiches had a higher rate of

energy absorption, the IPC30 sandwiches actually had greater energy absorption capacity

under impact conditions. When compared to IPC30 sandwiches, the SFSb30 sandwiches

had 22.1% less strain energy absorption and a 14.2% drop in specific energy absorption.

Both architectures experienced visible core crack initiation at ~550 µs. Note the

difference in Figures 6.43 and 6.44 after that time. SFS-b sandwich energy absorption

only occurs from time zero up to the time of crack initiation. The IPC foam core

sandwiches continue to absorb energy even after the onset of a crack.

The impact tests in this work were only conducted using 30% volume fraction of

microballoons in the SF. The purpose of these tests was to further expound on the failure

mechanisms in the quasi-static case, as well as begin to understand the failure

mechanisms in the dynamic case. Also, the feasibility of using DIC method in

conjunction with a high speed digital camera to study impact behavior of sandwich

structures is illustrated. Contrary to the quasi-static case, IPC foam core sandwiches

appear to outperform the syntactic foam core sandwiches with graded face sheets. The

presence of the interpenetrating phases seems to have a positive effect under dynamic

conditions, though the crack still initiates in the tensile region of the core.

143
CHAPTER 7

CONCLUSIONS

7.1 Comparison of Results and Conclusions

In this thesis, flexural responses of sandwich structures with syntactic foam (20%-

40% volume fractions) core were studied. Three types of sandwich structures were

studied. The first type had a regular syntactic foam core (identified as SFS) with simply

adhered aluminum face sheets. The second one employed an aluminum-syntactic foam

interpenetrating core (identified as IPC) in an attempt to prevent the face-core debond

seen in the SFS sandwiches. The third type consisted of a syntactic foam core with

graded face sheet (identified as SFS-b). Static three-point bend tests were carried out on

all three types of sandwich structures and global load-deflection responses were

measured. The three different architectures were comparatively examined in terms of

peak load, mid-span deflection at failure, nonlinearity of the response, and strain energy

absorbed. The global measurements were supplemented by digital image correlation

measurements on the core to map 2D deformations and strains. The global

measurements, measured normal and shear strain fields, and optical microscopy were

used to discern failure responses of the three architectures.

144
In the case of the syntactic foam core sandwiches (SFS), the load-deflection

behavior was essentially linear elastic in nature. However, for the IPC foam core

sandwiches and syntactic foam core sandwiches with graded face sheets (SFS-b), there is

a noticeable kink or a “knee”, giving the appearance of a bilinear flexural response.

The SFS sandwiches failed predominantly due to face-core debonding with a

large scatter in measured peak load and crosshead deflection at failure. They also

showed very limited nonlinearity in load-deflection response. The IPC core sandwiches

showed a substantial improvement in terms of deflection at failure and a significant

nonlinearity in load-deflection responses. The peak load, however, showed a slight

reduction. The nonlinearity in this architecture is primarily due to face sheet yielding as

well as debonding between the phases of the interpenetrating composite foam core. The

SFS-b sandwiches, on the other hand, showed substantial improvement in both the peak

load and mid-span deflection at failure and exhibited nonlinearity in the response similar

to IPC foam core sandwiches. This is attributed to a gradual shear strain variation at the

graded face-core interface (unlike the SFS sandwiches) resulting in face sheet yielding

before final failure.

Considering that the SFS material possibly did not reach its full potential and

comparing the specific strain energies (which accounts for the weight difference in

adding the aluminum foam into the core), IPC20 had a 25% improvement, IPC30 had a

59% improvement, and IPC40 had an 88% improvement. While the peak load decreased

slightly, for the small penalty of added weight the IPC foam core significantly improved

the value of deflection-to-failure experienced and thus the strain energy absorbed.

145
With the IPC30 having the highest average values for all metrics in the IPC foam

core test group (most pronounced specific energy absorbed), an optimum IPC foam

volume fraction of SF likely exists somewhere around 30%.

Though IPC foam core sandwiches outperformed the SFS counterparts, the

syntactic foam core sandwiches with graded face sheets (SFS-b) outperformed the IPC

foam core sandwiches. This architecture was able to maintain linear load-deflection

behavior much longer than the IPC foam core architecture with an improvement in load

at the (linear to nonlinear) transition point of 244%, 230%, and 211% for SFSb20,

SFSb30, and SFSb40, respectively. The improvements in mid-span deflection at the

transition point were 222%, 234%, and 225% for SFSb20, SFSb30, and SFSb40,

respectively. Comparing the specific strain energies to those of IPC foam core

sandwiches (which accounts for the weight difference in the core), SFSb20 had a 71%

improvement, SFSb30 had a 76% improvement, and SFSb40 had a 118% improvement.

Remember that ideally the data needs to be normalized by a sandwich consisting of

aluminum face sheets separated by an empty core. However, this being impractical, the

unfilled 20ppi aluminum foam core sandwich response was used to normalize the data.

Comparisons of all three architectures are shown in Figures 7.1 to 7.6.

146
25
SFS
IPC
SFSb
20

Peak Load/ALS Plateau Load 15

10

0
20% 30% 40%
Volume Fraction

Figure 7.1 Normalized Measured Peak Load for Sandwich Structures with SF and IPC

Foam Cores
14
SFS

12 IPC
SFSb
10
δFailure/δALS, nonlinear

0
20% 30% 40%
Volume Fraction

Figure 7.2 Normalized Measured Mid-Point (Load-Point) Deflection at Failure for

Sandwich Structures with SF and IPC Foam Cores

147
18
IPC
16
SFSb

Transition Load/ALS Plateau Load


14

12

10

0
20% 30% 40%
Volume Fraction

Figure 7.3 Normalized Measured Load at Transition Point for Sandwich Structures with

SF and IPC Foam Cores

6
IPC

5 SFSb

4
δTransition/δALS, nonlinear

0
20% 30% 40%
Volume Fraction

Figure 7.4 Normalized Measured Mid-Point (Load-Point) Deflection at Transition Point

for Sandwich Structures with SF and IPC Foam Cores

148
14
SFS
IPC
12 SFSb

10

Energy Absorbed (J) 8

0
20% 30% 40%
Volume Fraction

Figure 7.5 Energy Absorbed for Sandwich Structures with SF and IPC Foam Cores

200
SFS
180 IPC
SFSb
160
Specific Energy Absorbed (J/kg)

140

120

100

80

60

40

20

0
20% 30% 40%
Volume Fraction

Figure 7.6 Specific Energy Absorbed for Sandwich Structures with SF and IPC Foam

Cores

149
The superior material based on all these metrics is the syntactic foam core

sandwiches with graded face sheets (SFS-b). The only instance it might not be

considered the optimal choice might be when considering the IPC foam core sandwiches

for their ability to use the metallic foam network to hold it together longer. Depending

on the combination of desired properties, the volume fraction of microballoons could be

tailored to fit the needs of different applications.

Using the calculated strains from the image correlation process, plots of dominant

strain (εx) and shear strain (εxy) as a function of thickness were obtained. The strains

extracted were at x=25 millimeters (from the center of the beam). This distance from the

center was approximately half way between the loading and support rollers, minimizing

the stress concentration effects created by the rollers. The syntactic foam (SFS), IPC

foam, and syntactic foam (SFS-b) core material of the same volume fraction were

compared for a similar applied load. Figure 7.7 shows an example of the dominant

flexural normal strain εx results at a load of ~4,000 N for the 20% volume fraction of

microballoons in the SF.

150
Material: IPC/SFS
Temperature: 70°F
9.35 Rate: 0.025 in/min
Load: 4,000 N

6.85

4.35

1.85
y (mm)

FLEX-38
(SFS20)
-0.65 FLEX-10
(IPC20)
FLEX-57
-3.15 (SFSb20)

-5.65

-8.15

-10.65
-0.45 -0.35 -0.25 -0.15 -0.05 0.05 0.15 0.25 0.35 0.45
ε x (%)

Figure 7.7 20% Vf: Strain (εx) Variation Along the Height of the Sandwich Core

At 4,000N load, the SFS sandwich was still in the linear region, while the IPC

foam core and SFS-b sandwiches had entered the nonlinear regime. Notice that the SFS

and SFS-b sandwiches are generally exhibiting similar trends. This makes intuitive sense

considering the core material in each is the same syntactic foam. When comparing the

results with linear beam theory (where a linear strain variation through the thickness is

expected), one can observe that near the top and bottom of the beam, the behavior

deviates from linearity. This can be explained by the effects of face sheets restricting the

motion of the neighboring material points in the core, and by the increased amount of

image processing errors near the face sheets. Also, there is a shift in the neutral axis from

the center of the beam in the SF core cases. This can be explained by the severity of the

stress concentration effects of the loading point in the SF core cases. The expected

general trend of greater strain experienced in the IPC foam core compared to syntactic

foam of the same volume fraction is confirmed. Remember that the SFS sandwiches

151
failed because of the intense shear stresses seen at the faces, causing face-core

debonding. The main idea to take away from the shear strain plots is that both IPC core

and SFS-b sandwiches have done a better job of mitigating the intensity of the shear

strains (and thus stresses) at the faces. Figure 7.8 shows an example of the εxy results at

a load of ~4,000N for the 20% volume fraction of microballoons in the SF.

Material: IPC/SFS
Temperature: 70°F
9.35 Rate: 0.025 in/min
Load: 4,000 N

6.85

4.35

1.85
y (mm)

FLEX-38
(SFS20)
-0.65 FLEX-10
(IPC20)
FLEX-57
-3.15 (SFSb20)

-5.65

-8.15

-10.65
0.0000 0.0005 0.0010 0.0015 0.0020 0.0025 0.0030 0.0035 0.0040
ε xy (rad)

Figure 7.8 20% Vf: Shear Strain (εxy) Variation Along the Height of the Sandwich Core

In the impact tests, two major aspects of note are the rate of energy absorption and

the total strain energy absorbed by the specimen. While the SFSb30 sandwiches had a

higher rate of energy absorption, the IPC30 sandwiches actually had greater energy

absorption capacity under impact conditions. When compared to IPC30 sandwiches, the

SFSb30 sandwiches had 22% less strain energy absorption and a 14% drop in specific

energy absorption. The SFSb30 absorbed energy at a rate of ~20 J/ms, while the IPC30

152
absorbed at a rate of ~12 J/ms. Both architectures experienced visible core crack

initiation around the same time. SFS-b sandwich energy absorption only occurs from

time zero (impact) up to the time of crack initiation. The IPC foam core sandwiches,

however, continue to absorb energy even after the onset of a crack resulting in greater net

energy absorbed until complete failure.

7.2 Future Work

The primary focus of this work was to understand flexural failure behavior of IPC

and syntactic foam core sandwiches under quasi-static loading conditions. An attempt

was made to illustrate the feasibility of studying the same under impact loading

conditions. Experiments indicate an improved response of the IPC foam core sandwich

over the SFS sandwich, including energy absorption. However, in light of the failure

mechanism typically observed in the IPC foam core, an effort was made to create a

syntactic foam core sandwich with a graded face sheet to promote better load transfer

throughout the sandwich. The superior material that emerged and is recommended is the

SFS-b sandwich architecture.

Since most structures undergo combined loading it would be of interest to fully

characterize the core material in tension and bending. This study has not tested any of

the core material without face sheets, and therefore cannot speculate on the properties of

the core alone. One particular future objective of interest would be to obtain mechanical

properties (elastic modulus, Poisson’s ratio) of the core materials in bending and tension.

Jhaver [17] was able to obtain mechanical properties of the materials in question, but in

153
compression only. While the experimentally derived properties from compression should

be close to those seen in bending, it would be advantageous to confirm this

experimentally.

The failures seen in the SFS sandwiches stemmed from debonding of the face-

core interface. The face sheet would typically separate from the core, leaving the core as

the only load bearing component. Very shortly after this debond, the core would

catastrophically fail. This is unlike the IPC sandwich where a more synergistic

relationship between the core and face was observed. However, with the IPC core

sandwich, interphase separation (between aluminum ligaments and SF) in the tensile

region was the source of failure. To better harness the face-core relationship for the

syntactic foam core, smoother transition from face sheet to core needs to exist. One

possible solution is to have face sheets with a thin layer of brazed aluminum foam similar

to that of the IPC. A better response was achieved by better linking the face and core

(SFS-b sandwiches), preventing the debond typically seen in the SFS sandwiches. The

useful extent of this thin layer of foam was not explored here. In this work all of the

graded face sheets had a layer of ~2 mm of aluminum foam brazed to the aluminum face

sheet. It would be beneficial to test varying thicknesses of this brazed layer of foam to

determine at what depth the best overall response occurs while still maintaining the

integrity of the syntactic foam core. Another closely related idea that could be explored

would be to change the method by which the core and face are fused. With such stiff

materials currently being used, it might be advantageous to attempt to use a more

compliant layer as a transition layer from the core to the faces.

154
The impact tests in this work were only conducted using 30% volume fraction of

microballoons in the SF. The purpose of these tests was to further elaborate on the

failure mechanisms in the quasi-static case, as well as begin to understand the failure

mechanisms in the dynamic case. Future work should involve varying the microballoon

volume fraction. Also, the feasibility of using DIC method in conjunction with a high

speed digital camera to study impact behavior of sandwich structures is illustrated.

Contrary to the quasi-static case, IPC foam core sandwiches appear to outperform the

syntactic foam core sandwiches with graded face sheets. The presence of the

interpenetrating phases seems to have a positive effect under dynamic conditions. These

tests serve as a first step on the path to dynamic characterization of sandwich structures

with syntactic foam and IPC foam core architectures.

155
REFERENCES

1. Gibson, L.J., Ashby, M.F., 1988. Cellular Solids: Structure and Properties.
Pergamon Press, Oxford, England.

2. Wu, C.L., Weeks, C.A., Sun, C.T. Improving Honeycomb-Core Sandwich Structures
for Impact Resistance. Journal of Advanced Materials 26 (1995) 41-47.

3. Zarei, H.R., Kröger, M. Bending Behavior of Empty and Foam-Filled Beams: An


Optimization. International Journal of Impact Engineering 35 (2008) 521–529.

4. Seitzberger, M., Rammerstorfer, F.G., Gradinger, R., Degischer, H.P., Blaimschein,


M., Walch, C. Experimental Studies on the Quasi-Static Axial Crushing of Steel
Columns Filled with Aluminum Foam. International Journal of Solids and Structures
37 (2000) 4125-4147.

5. Belingardi, G., Cavatorta, M.P., Duella, R. Material Characterization of a Composite-


Foam Sandwich for the Front Structure of a High Speed Train. Composite Structures
61 (2003) 13–25.

6. Zhang, Q., Lee, P.D., Singh, R. Wu, G., Lindley, T.C. Micro-CT Characterization of
Structural Features and Deformation Behavior of Fly Ash/Aluminum Syntactic Foam.
Acta Materialia 57 (2009) 3003–3011.

7. Zhou, W., Hu, W., Zhang, D. Study on the Making of Metal-Matrix Interpenetrating
Phase Composites. Scripta Materialia, Vol. 39, No. 12 (1998) 1743–1748.

8. Styles, M., Compston, P., Kalyanasundaram, S. Flexural Behaviour of


Aluminum/Foam Composite Structures: An Investigation into the Deformation
Mechanism and Strain Distribution under 4 Point Bending in Comparison with
Polymer Foam Core Structures. Sandwich Structures 7: Advancing with Sandwich
Structures and Materials (2005) 487–496.

156
9. Styles, M., Compston, P., Kalyanasundaram, S. The Effect of Core Thickness on the
Flexural Behaviour of Aluminum Foam Sandwich Structures. Composite Structures
80 (2007) 532–538.

10. McKown, S., Mines, R.A.W. Static and Impact Behaviour of Metal Foam Cored
Sandwich Beams. High Performance Structures and Materials II (2004) 37-46.

11. McCormack, T.M., Miller, R., Kesler, O., Gibson, L.J. Failure of Sandwich Beams
with Metallic Foam Cores. International Journal of Solids and Structures 38 (2001)
4901-4920.

12. Chen, C., Harte, A-M., Fleck, N.A. The Plastic Collapse of Sandwich Beams with a
Metallic Foam Core. International Journal of Mechanical Sciences 43 (2001) 1483-
506.

13. Dukhan, N., Rayess, N., Hadley, J. Characterization of Aluminum Foam-


Polypropylene Interpenetrating Phase Composites: Flexural Test Results. Mechanics
of Materials 42 (2010) 134–141.

14. Travitzky, N.A. Effect of Metal Volume Fraction on the Mechanical Properties of
Alumina/Aluminum Composites. Journal of Materials Science 36 (2001) 4459–4463.

15. Chang, H., Higginson, R., Binner, J. Microstructure and Property Characterisation of
3-3 Al(Mg)/Al2O3 Interpenetrating Composites Produced by a Pressureless Infiltration
Technique. J Mater Sci 45 (2010) 662–668.

16. Allen, H.G., 1969. Analysis and Design of Structural Sandwich Panels. Pergamon
Press, Oxford, England.

17. Jhaver, R. Compression response and modeling of interpenetrating phase composites


and foam-filled honeycombs. M.S. Thesis, Auburn University, 2009.

18. Sadler, E.J., Vecere, A.C. Silane Treatment of Mineral Fillers – Practical Aspects.
Plastics, Rubber, and Composites Processing and Applications 24 (1995) 271-275.

19. ASTM C393/C393M-06. Standard Test Method for Core Shear Properties of
Sandwich Constructions by Beam Flexure. 2006.

20. Kobayashi, A.L., 1987. Handbook on Experimental Mechanics. Prentice-Hall, Inc.,


Englewood Cliffs, New Jersey, USA.

21. Chu, T.C., Ranson, W.E., Sutton, M.A., Peters, W.E. Application of Digital Image
Correlation Techniques to Experimental Mechanics. Experimental Mechanics 25 (3)
(1985) 232-244.

157
22. McNeill, S.R., Sutton, M.A., Miao, Z., Ma, J. Measurement of Surface Profile Using
Digital Image Correlation. Experimental Mechanics 37 (1) (1997) 13-20.

23. Sun, Z., Lyons, J.S., McNeill, S.R. Measuring Microscopic Deformations Using
Digital Image Correlation. Optics and Lasers in Engineering 27 (1997) 409-428.

24. Tyson, J., Schmidt, T., Galanulis, K. Advanced Photogrammetry for Robust
Deformation and Strain Measurement. Proceedings of SEM 2002 Annual
Conference, Milwaukee, WI, June 2002.

25. Tyson, J., Schmidt, T., Galanulis, K. Dynamic Strain Measurement Using Advanced
3D Photogrammetry. Proceedings of IMAC XXI, Kissimmee, FL, Feb 2003.

26. ARAMIS™ User Manual (Version 6.1). GOM mbH, Germany.

158
APPENDIX A
GLOBAL/LOCAL DISPLACEMENT COMPARISONS

The following are comparisons between the global load-displacement data from
the machine and the local displacement data from the image correlation for each of the
specimens used as examples throughout the thesis.

6000 Material: SFS20


Temperature: 70°F
Specimen: FLEX-38
Cast: 12Jul2010
Test: 14Sept2010
5000 Rate: 0.025 in/min

4000
Load (N)

3000 Crosshead

ARAMIS

2000

1000

0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Displacement (mm)

Figure A.1 Global/Local Displacement Comparison, FLEX-38 (SFS20)

159
6000 Material: SFS30
Temperature: 70°F
Specimen: FLEX-39
Cast: 04Oct2010
Test: 26Oct2010
5000 Rate: 0.025 in/min

Load (N) 4000

3000 Crosshead

ARAMIS

2000

1000

0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
Displacement (mm)

Figure A.2 Global/Local Displacement Comparison, FLEX-39 (SFS30)

4500 Material: SFS40


Temperature: 70°F
Specimen: FLEX-24
Cast: 18May2010
4000 Test: 15Jul2010
Rate: 0.025 in/min

3500

3000

2500
Load (N)

Crosshead

2000

ARAMIS

1500

1000

500

0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Displacement (mm)

Figure A.3 Global/Local Displacement Comparison, FLEX-24 (SFS40)

160
6000 Material: IPC20
Temperature: 70°F
Specimen: FLEX-10
Cast: 30Sept2009
Test: 21Apr2010
5000 Rate: 0.025 in/min

Load (N) 4000

3000 Crosshead

ARAMIS

2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5
Displacement (mm)

Figure A.4 Global/Local Displacement Comparison, FLEX-10 (IPC20)

6000 Material: IPC30


Temperature: 70°F
Specimen: FLEX-16
Cast: 15Apr2010
Test: 09Jun2010
Rate: 0.025 in/min
5000

4000
Load (N)

3000 Crosshead

ARAMIS

2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Displacement (mm)

Figure A.5 Global/Local Displacement Comparison, FLEX-16 (IPC30)

161
5000 Material: IPC40
Temperature: 70°F
Specimen: FLEX-14
Cast: 07Oct2009
Test: 21Apr2010
Rate: 0.025 in/min
4000

3000
Load (N)

Crosshead

2000 ARAMIS

1000

0
0.0 0.5 1.0 1.5 2.0 2.5
Displacement (mm)

Figure A.6 Global/Local Displacement Comparison, FLEX-14 (IPC40)

9000 Material: SFSb20


Temperature: 70°F
Specimen: FLEX-57
Cast: 23Jan2011
8000 Test: 10Feb2011
Rate: 0.025 in/min

7000

6000

5000
Load (N)

Crosshead

4000

ARAMIS

3000

2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5
Displacement (mm)
Figure 1.1-1 Flexural Response of SF Sandwich (20%) at Room Temperature

Figure A.7 Global/Local Displacement Comparison, FLEX-57 (SFSb20)

162
9000 Material: SFSb30
Temperature: 70°F
Specimen: FLEX-59
Cast: 25Jan2011
8000 Test: 02Feb2011
Rate: 0.025 in/min

7000

6000

5000
Load (N)

Crosshead

4000

ARAMIS

3000

2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Displacement (mm)
Figure 1.1-1 Flexural Response of SF Sandwich (30%) at Room Temperature

Figure A.8 Global/Local Displacement Comparison, FLEX-59 (SFSb30)

8000 Material: SFSb40


Temperature: 70°F
Specimen: FLEX-65
Cast: 01Feb2011
7000 Test: 10Feb2011
Rate: 0.025 in/min

6000

5000
Load (N)

4000 Crosshead

ARAMIS
3000

2000

1000

0
0.0 0.5 1.0 1.5 2.0 2.5
Displacement (mm)
Figure 1.1-1 Flexural Response of SF Sandwich (40%) at Room Temperature

Figure A.9 Global/Local Displacement Comparison, FLEX-65 (SFSb40)

163
APPENDIX B
SELECT ARAMIS™ IMAGES

The series of images shown in Figure B.1 are typical ARAMIS™ εy results for

tests conducted on IPC sandwiches. This case (IPC30) is simply used as an example to

illustrate the εy trends. In the set of four images, the first image corresponds to a point in

the linear regime, the second image corresponds to a point in the nonlinear regime, the

third image corresponds to the point just before failure, and the fourth image corresponds

to the point just after failure. Note the units are measured in percent. Also note that the

correlated images for IPC30 contain twenty contours with an increment of 0.15%. The

strains are evenly distributed, with muted concentration effects.

The series of images shown in Figure B.2 are typical ARAMIS™ εy results for

tests conducted on SFS-b sandwiches. Again, this case (SFSb30) is used as an example.

In the set of three images, the first image corresponds to a point in the linear regime, the

second image corresponds to a point in the nonlinear regime, and the third image

corresponds to the point just before failure. As before, the units are in percent. The

correlated images for SFSb30 also contain twenty contours, but this time with an

increment of 0.5%. Notice the concentration effects seen in this case.

164
y

x P=1097N

P=3997N

P=5680N

P=3741N

Figure B.1 ARAMIS™ Results for IPC30 (FLEX-16, εy)

165
y

x P=2990N

P=6717N

P=7559N

Figure B.2 ARAMIS™ Results for SFSb30 (FLEX-59, εy)

166

Potrebbero piacerti anche