You are on page 1of 152

DEVELOPING CONSTITUTIVE EQUATIONS FOR POLYMER FOAMS

UNDER CYCLIC LOADING

A Dissertation

Presented to

The Graduate Faculty of The University of Akron

In Partial Fulfillment

of the Requirements for the Degree

Doctor of Philosophy

Linling Chen

December, 2012
DEVELOPING CONSTITUTIVE EQUATIONS FOR POLYMER FOAMS

UNDER CYCLIC LOADING

Linling Chen

Dissertation

Approved: Accepted:

____________________________ ____________________________
Advisor Department Chair
Dr. Michelle S. Hoo Fatt Dr. Celal Batur

_____________________________ ____________________________
Committee Member Dean of the College
Dr. Atef Saleeb Dr. George K. Haritos

_____________________________ ____________________________
Committee Member Dean of the Graduate School
Dr. Gregory Morscher Dr. George R. Newkome

_____________________________ ____________________________
Committee Member Date
Dr. Xiaosheng Gao

_____________________________
Committee Member
Dr. Mukerrem Cakmak

ii
ABSTRACT

Sandwich constructions with polymer foam cores have widespread applications

by offering high bending stiffness and strength per unit weight. They can also dissipate

energy when subjected to shock and impact loadings. The current foam models are based

on metallic foams, which can describe polymer foam behavior under monotonic loading

but not under cyclic or vibration loadings. Cyclic compression and pure shear tests were

done on Divinycell PVC H100 foam to produce hysteresis data in compression and shear

in the in-plane and out-of-plane directions of the foam with strain rates ranging from

0.0005 s-1 to 5 s-1. The compressive strain amplitudes were limited to less than 10%,

while the engineering shear strain amplitudes were less than 20%.

Two one-dimensional constitutive models were developed from the test data. The

first constitutive model consisted of an equilibrium spring in parallel with a Maxwell

element and in series a Prandtl element. Material properties were introduced for

permanent crushing of cells (plastic deformation) and progressive damage of cells

(Mullins effect). This constitutive model was able to capture yielding, viscoelastic and

viscoplastic response, as well as hysteresis of the foam, but it failed to describe the strain-

rate dependency that was found in the experiments. The second constitutive model

consisted of an equilibrium spring only in parallel with a Maxwell element. This model

iii
was based on the concept of damage initiation and evolution of the foam. The damaged

material properties were found to depend on the magnitude and history of the

deformation. Good agreement with out-of-plane compression test results was reported

with the second constitutive model.

iv
ACKNOWLEDGEMENTS

This study was funded in part by ONR Grant No. N00014-11-1-0485. I would

like to acknowledge the kind support of Dr. Yapa D. S. Rajapakse, Program Manager of

Solid Mechanics Program at the Office of Naval Research.

The greatest thanks to my advisor, Dr. Hoo Fatt, from whom, I learned how to

conduct research that is needed to meet the high standards in the field. Her continuous

encouragement, enthusiasm and determination, as well as her outstanding insights and

dedication to her work have always inspired me. I would like to thank her for the pleasant

and stimulating discussions, and the time she spent in guiding me throughout my

graduate studies.

I would like to also express my sincere appreciation to all the committee

members: Dr. Saleeb for his great and most helpful advice on material modeling and

numerical implementation; Dr. Morscher for his kindness in answering my questions

about composites and his generosity of allowing me using his lab equipment; Dr. Gao for

continuum mechanics advice; Dr. Cakmak for his suggestions of the importance in

exploring the micro-structure of polymers. I want to thank you all for taking the time in
v
reading this dissertation. Your useful suggestions and constructive comments have

significantly improved the quality of this work.

I am also grateful to the Department of Mechanical Engineering for the financial

support throughout my doctoral program. I would like to thank Mr. Dale Ertley, together

with who we designed test apparatus and achieved an invention disclosure. I would like

also thank to Mr. Stephen Gerbetz, Mr. Clifford Bailey, Mrs. Christina L. Christian and

Mrs. Stacy Meier for their technical supports and personal care.

Special thanks to my parents, for their love, patience, encouragement,

understanding and sacrifices throughout my doctoral program.

Linling CHEN

Akron, December 2012

vi
TABLE OF CONTENTS

Page

LIST OF TABLES……………………………………………………………………...…x

LIST OF FIGURES………………………………………………………………………xi

CHAPTER

I. INTRODUCTION……………………………………………………………………....1

II. LITERATURE REVIEW………………………………………………………………7

2.1 Experimental Investigation………………………...………………………...…….8

2.2 Finite Element Models for Foam…………………………………………………10

2.2.1 Quasi-static and low-velocity impact…………………………….……...….11

2.2.2 High-velocity impact………………………………………………..……...12

2.2.3 Cyclic loading……………………………………………………………....13

2.3 Constitutive Models for Foam……………………………………………..…….16

2.3.1 Crushable foam model………………………………………………...……16

2.3.2 Viscoelasticity………………………………………………………...…….25

2.4 Energy Dissipation under Cyclic Loading………………………………..……...27

2.5 Combination of the Existing Theories………………………………..………….30

III. EXPERIMEMTS…………………………………………………………………….32

3.1 Specimens…………………………………………….………………………..…33

3.2 Test Plan………………….…………………………………………………..…..35

xv
3.3 Compressive Tests……………………………………………...…………..…...37

3.3.1 Apparatus design and test setup……………………………………..……..37

3.3.2 Monotonic tests results…………………………..………………………...40

3.3.3 Cyclic compression tests results……………………..…………………….46

3.3.4 Material homogeneity…………………………………..……………....….52

3.4 Pure shear tests……………………………………………………..………..…...61

3.4.1 Apparatus design and tests setup………………………………...………...61

3.4.2 Pure shear test results………………………………………….……..…….63

3.5 Summary and Conclusions………………………………………….….………..70

IV. CONSTITUTIVE MODELING………………………………...………….………..71

4.1 Damage with Mullins Effect…………………………………………..…....…...71

4.1.1 Viscoelasticity…………………………………………………….…....….73

4.1.2 Viscoplasticity…………………………….……………………………….81

4.1.3 Mullins Effect………………………….………………………………….88

4.2 Damage Mechanics Model…………………….………………………………..97

V. CONCLUSIONS AND FUTURE WORK…………….……………………….…...106

REFERENCES……………………………………………………….……………...…108

APPENDICES …..……………………………………………….….…………..….….114

APPENDIX A UNIAXIAL COMPRESSION TEST RESULTS……………………...114

APPENDIX B CYCLIC COMPRESSION TEST RESULTS…………………………116

APPENDIX C CYCLIC SHEAR TEST RESULTS…………………………………...121

APPENDIX D COMPARISON OF VISCOELASTIC CURVES……………………..126

APPENDIX E COMPARISON OF VISCOPLASTIC CURVES……………………...130

viii
APPENDIX F INELASTIC STRAIN VALUES……………………………………….132

APPENDIX G COMPARISON OF PREDICTED RESPONSE TO TEST RESULTS


FOR OUT-OF-PLANE COMPRESSION USING VISCOELASTIC DAMAGE MEDEL
……………………………………………………………………………………..……133

ix
LIST OF TABLES

Table Page

3.1 Characteristics Divinycell H100 [2]………………………………………….….…..33

3.2 Test frequency of MTS machine in cyclic compression tests…………………...…...36

3.3 Test frequency of MTS machine in pure shear tests…………………………..….….37

3.4 Out-of-plane strain rate effect for PVC H100………………………………….…….43

4.1 Material properties of out-of-plane and in-plane after curve fitting……………........78

4.2 Constant material properties of out-of-plane and in-plane……………………..…....79

4.3 Comparison of results for viscoelasticity in out-of-plane and in-plane


directions……………………………………..……………………...……..………...81

4.4 Material properties for viscoplasticity……………………………….……………....86

4.5 Comparison results for viscoplasticity in out-of-plane and in-plane


directions…………………………………...…………………………….……...…...88

4.6 Flow stresses with various strain rates in out-of-plane compression...……….….....100

x
LIST OF FIGURES

Figure Page

2.1 A typical foam compression stress-strain curve [25]…………..…..………………….8

2.2 Projectile impact into sandwich panel:


(a) FEA results and (b) Hysteresis in the foam………………..…………………..…14

2.3 Blast loading of a sandwich shell:


(a) FEA model and (b) Hysteresis in the foam………………………….…………...15

2.4 Crushable foam model with isotropic hardening:


yield surface and flow potential in the p–q stress plane [11]……….………………..19

2.5 Crushable foam model with volumetric hardening:


yield surface and flow potential in the p–q stress plane [11]…….…………………..19

2.6 Stress-strain responses of crushable foam models:


(a) Isotropic hardening and (b) Volumetric hardening……………………………....21

2.7 Bauschinger effect…………………………………………………………………. 22

2.8 Modulus reduction:


(a) Shear stress-strain ( and (b) Shear modulus ratio verses shear strain……..23

2.9 Modulus degradation:


(a) Shear stress-strain ( and (b) Shear modulus ratio verses number of cycles.24

2.10 Effect of strain rate on the compressive behavior:


(a) Idealized metallic foam and (b) Divinycell PVC H200 foam [38]……………..25

2.11 Three viscoelastic models:


(a) Maxwell; (b) Voigt and (c) Linear Spring-Maxwell model…………………....26

2.12 Schematic loading-unloading curves in simple tension [45]……………………….28

3.1. Divinycell PVC H100 material sheet………………………………………….…….34

3.2 Strain history at a strain rate 5 s-1 and strain amplitude 4%.........................................36

xi
3.3 PVC H100 glued on Aluminum gripes………………………………………………39

3.4 Set up for compressive cyclic test for PVC H100…………………………………...39

3.5 Set up of “C” clamps:


(a) A pair of “C” clamps and (b) Side views of “C” clamp………………………….40

3.6 Procedure to define linear portion in the elastic region……………………………...41

3.7 Out-of-plane monotonic test under strain rate of 0.1 s-1……………………………..42

3.8 Out-of-plane monotonic test at various strain rates………………………………….42

3.9 Out-of-plane vs. in-plane at strain rate of 0.1 s-1………………………………….…44

3.10 Out-of-plane compressive stress-strain curves for glued and unglued specimens
at strain rate of 0.1s-1………………………………………………………………..45

3.11 Out-of-plane tensile test vs. compressive test………………………………………46

3.12 Out-of-plane compressive stress-strain curves


at strain amplitude of 0.10 with various strain rates………………………………..47

3.13 In-plane compressive stress-strain curves


at strain amplitude of 0.10 with various strain rates………………………………..48

3.14 Out-of-plane compressive stress-strain curves for various cycles………………….48

3.15 Out-of-plane compressive stress-strain curves


at strain rate of 0.5 s-1 with various strain amplitudes……………………………...49

3.16 Compressive stress-strain curve at a strain rate of 0.5 s-1 with various
strain amplitudes: (a) Out-of-plane and (b) In-plane………………………......…..50

3.17 Out-of-plane first cyclic of strain rates of 0.5 s-1 with various
strain amplitudes over lamping on the primary curve……………………………...51

3.18 Out-of-plane fourth cycle hysteresis at 10% strain at various strain rates………….51

3.19 PVC H100 crushed with 10% strain amplitude at strain rate of 0.0005 s-1
in the in-plane direction: (a) Crushed specimen (b) Undamaged micro-structure at
A2 and (c) Damaged micro-structures at various locations……….………………...54

3.20 PVC H100 crushed with 10% strain amplitude at strain rate of 0.0005 s-1
in the out-of-plane direction: (a) Crushed specimen (b) Undamaged micro-structure
at A2 and (c) Damaged micro-structures at various locations.…………………….55

xii
3.21 PVC H100 crushed with 10% strain amplitude at strain rate of 5 s-1 in the in-plane
direction: (a) Crushed specimen (b) Undamaged micro-structure at A2 and
(c) Damaged micro-structures at various location…………….…………………56

3.22 PVC H100 crushed with 10% strain amplitude at strain rate of 5 s-1
in the out-of-plane direction: (a) Crushed specimen (b) Undamaged micro-structure
at A2 and (c) Damaged micro-structures at various locations…………………...…57

3.23 Crushed specimens loaded to 10% strain amplitude in the in-plane direction with
five increasing strain rates………………………………………………………….58

3.24 Crushed specimens loaded to 10% strain amplitude in the out-of-plane direction
with five increasing strain rates…………………………………………………….58

3.25 PVC H100 crushed with 30% strain amplitude at strain rate of 0.0005 s-1
in the in-plane direction: (a) Crushed specimen (b) Undamaged micro-structure at
A2 and (c) Damaged micro-structures at various locations……..……….…………59

3.26 PVC H100 crushed with 30% strain amplitude at strain rate of 0.0005 s-1
in the out-of-plane direction: (a) Crushed specimen (b) Undamaged micro-structure
at A2 and (c) Damaged micro-structures at various locations……………….…….60

3.27 Pure shear test apparatus:


(a) Schematic of shear text fixture and
(b) Photograph showing experimental setup………………….....………………….62

3.28 Comparison of out-of-plane shear stress-strain curves of PVC H100 foam core at
strain rate =0.001 s-1 with data from Steeves and Fleck (2004) and Danel (2009)…65

3.29 Out-of-plane shear fracture of PVC H100 foam core at strain rate =0.001 s-1……..66

3.30 Monotonic shear test at strain rate=0.5 s-1, strain amplitude=0.20………………....67

3.31 Pure shear stress-strain curves of strain amplitude of 20% at various strain rates:
(a) Out-of-plane and (b) In-plane……………………………………………...……68

3.32 Pure shear stress-strain curves at a strain rate of 0.5 s-1 with various strain
amplitudes: (a) Out-of-plane and (b) In-plane……………………………...………69

4.1 Material behavior of strain rate of 0.5 s-1 (out-of-plane)…………………………….72

4.2 Mechanical analogs model…………………………………………………………...72

4.3 Experimental stress-strain curves showing viscoelasticity:


(a) Out-of-plane and (b) In-plane…………………………………………………….74

xiii
4.4 Mechanical analogs for a viscoelastic………………………………..………………75

4.5 Relationship of strain verses time for loading, unloading, and reloading………..….77

4.6 Fitted curve for viscoelastic response:


(a) Out-of-plane and (b) In-plane…………………………………………………….80

4.7 Elastic-plastic stress-strain curve in metals…………………………...………….….81

4.8 Plastic hardening (a) Isotropic hardening and (b) Kinematic hardening………...….82

4.9 Experimental stress-strain curves showing viscoplasticity:


(a) Out-of-plane and (b) In-plane……………………………………………...….….85

4.10 Fitted curves for viscoplastic response:


(a) Out-of-plane and (b) In-plane……………………...……………..……..………87

4.11 Stress-strain relations after plastic flow…………………………………………….88

4.12 Modified mechanical analogs model……………………………………………….90

4.13 Obtaining plastic strain damage function.………………………………………….92

4.14 The experimental curve verses the simulation of Mullins damage at strain rate of
0.0005 s-1 in the out-of-plane direction…………………………………………….93

4.15 Comparison of the hysteresis at different strain rates in out-of-plane direction with
r=1.29, m=0.035, C1=2 and C2=220. (a) 0.0005 s-1, (b) 0.005 s-1, (c) 0.05 s-1,
(d) 0.5 s-1and (e) 5 s-1…………………….………………………………………….95

4.16 Comparison of the hysteresis at different strain amplitude in out-of-plane direction


with r=1.29, m=0.035, C1=2 and C2=220. (a) 10%, (b) 8%, (c) 6% and (d) 4%. …96

4.17 Damage initiations at using actual test for out-of-plane direction...……….97

4.18 Stress-strain response using damage mechanics:


(a) History effects and (b) Effect of maximum strain amplitude……..………….....99

4.19 Determining plastic stress at , using simulated viscoelastic response…….100

4.20 Mechanical analog for a damage mechanics model:


(a) Before damage and (b) After damage…………………..………………………101

4.21 Comparison of experimental data with simulation curves at strain amplitude of 10%
in the out-off-plane direction: (a) 0.0005 s-1, (b) 0.005 s-1, (c) 0.05 s-1, (d) 0.5 s-1and
(d) 5 s-1…...…………………………………………………………………………104

xiv
4.22 The comparison at various strain rates of strain amplitude of 10% in the out-of-plane
direction………………………………………………………………………..….105

4.23 Comparison of simulation curve and test data at various strain amplitudes and fixed
strain rate of 0.0005 s-1 in the out-of-plane direction………………………….......105

xv
CHAPTER I

INTRODUCTION

Natural disasters and terrorists’ bomb blasts injure and kill people and destroy

property, causing billions of dollars in damages. How to prevent and limit the damage

due to them has become a big issue to both political leaders and scientists. As part of

these efforts, scientists are studying ways to improve the structures’ resistance to high

intensity dynamic loads.

The sandwich configuration is considered as one of the most efficient structural

designs not only because it offers high bending stiffness and strength per unit weight, but

it may dissipate energy when subjected to shock and impact loadings. In 1820, a

Frenchman, Duleau [1] first discussed the advantages of using two co-operating faces

separated by some distance to create a sandwich structure. It was not until 110 years later

that the concept was first applied commercially. During World Wars I and II, the use of

sandwich panels became widespread. The construction consisted of a lightweight core

sandwiched between two stiff facesheets. The sandwich material options were numerous,

such as metallic and composite facesheets with honeycomb, balsa wood and foam cores,

and a large range of adhesives. Reports [2] claims that the facesheets carry bending

1
stresses and give the structure a hardwearing surface, while the core material absorbs the

transverse shear stresses generated by loads and distribute them over a larger area.

For decades, sandwich constructions with foam cores have been utilized in

marine, civil, transportation and defense industries to lighten, stiffen and

strengthen components ranging from bridges, buildings, ships, trains, and even

windmill blades. The sandwich structures have high specific stiffness and

strength, noise reduction, thermal insulation, and impact energy absorption

characteristics. For instance, in the transportation industry, sandwich structures

have been used in high-speed truck trailers and high-speed trains for the safety of

passengers. Sandwich constructions also act as the panels mounted outside a

building that will be subjected to wind loads [3] or even terrorist’s bomb blasts.

On a daily basis, external structures experience rain, sleet, snow and hail; when

under terrorists’ attack; external sandwich panels become a shield to protect

people on the inside. Under bomb blast, not just the blast but fragments or debris

from walls, windows and fixtures, become high-velocity projectiles causing injury

and death. A sandwich panel can provide a nesting zone for these fragments.

Indeed, the study of sandwich structures that have both high static strength and

impact energy absorption capability is very important.

Foam-core panels are stronger than solid laminates, only if they are flat.

Curved foam-core panels are weaker than solid glass panels, especially when the

compressive load is applied in shear mode. Scientists, who analyze failures of

2
curved sandwich hulls in sailboats, found out that curved foam-core panels are poor in

bending [4]. Hence, the interest of this research is focused on foam-core flat panels with

various polymer foam core materials, such as polyvinylchloride (PVC). Nowadays, when

cost is the typical driving force in design, polymer foams are superior to other core

materials, such as metal honeycomb and balsa wood, because they are cheap and require

no maintenance. For example, PVC foam is half the price of aluminum honeycomb and

four times cheaper than balsa wood.

PVC foam has “closed cells,” which makes this material water-resistant and anti-

corrosive. It is also considered as a non-toxic pollutant (heavy metal-free materials) and

eco-friendly material [3]. For instance, Foamacell TM Green is recycled PVC foam made

with 70% recycled PVC and 30% virgin petrochemicals. Furthermore, PVC foam is

capable of cushioning forces and absorbing energy due to viscoelastic/viscoplastic cell

wall buckling, fracture, friction and viscous air/fluid flow. This ability of the polymer

foam is what makes it the ideal material for safety helmets, packaging and blast

protection. In these applications the foam is not bonded to two facesheets, as in typical

sandwich construction. It is used as a sacrificial element, to be permanently crushed or

destroyed. The same cushioning and energy absorption of the foam core, however, also

become important when sandwich panels are under dynamic loading such as impact or

pressure pulse (impulsive or blast loading). The difference in the case of a sandwich

panel is that the foam is bonded to two facesheets, and energy can be absorbed in the

form of kinetic and strain energy when the entire sandwich construction vibrates. The

foam in a sandwich structure dissipates energy through hysteresis from subsequent stress

3
wave propagation and structural vibrations. It is not intended to be a disposal part

of the system.

Previous work on sandwich panels subjected to impact loadings can be

divided in two regimes: low-velocity impact and high-velocity impact. Many

articles have been published on the behavior of sandwich structures subjected to

low-velocity impact, including the analysis of the influence of the foam core.

However, there is a lack of studies about their behavior under high-velocity

impacts [5]. According to the comprehensive review by Abrate [6], high-velocity

impact is a phenomenon controlled by wave propagation, and is essentially

independent of boundary conditions, whereas a low-velocity impact is highly

influenced by the boundary conditions. The study of high-velocity cases is

challenging.

There is no denial that experimental studies provide essential information,

especially since the impact phenomena depends on numerous parameters. A

comprehensive knowledge of the ballistic behavior of foams requires a broad test

program, which is time consuming and expensive. Therefore the use of analytical

[7] and numerical [8] models to analyze the perforation of sandwich structures is

critical to reduce cost and time in the design processes. The main advantage of

analytical models is the quick analysis of the influence of different parameters on

the high-velocity impact behavior of sandwich structures. Since the foam core

plays a major role in absovbing energy, the characteristics of the foam need to be

4
studied well. Experimental and analytical investigations of the crushing of foam are done

in this study.

There are two well-established constitutive models being used on polymer foams

today: crushable foam model [9-13] and hyperfoam model [14-17]. Both of these

constitutive models were initially proposed to address the monotonic compressive or

crushing response of the foam. Neither of them address viscoelasticity nor hysteresis,

although the hyperfoam model is often coupled with finite strain viscoelastic models to

produce hysteresis response. The hyperfoam model can also be combined with a Mullins

damage model to address stress softening during subsequent cyclic loads [18]. Several

approaches to crushing behavior can be found in literatures. More details will be

discussed in Chapter II. The crushable-foam plasticity model implemented in ABAQUS

has been validated with experimental results obtained from quasi-static and low-velocity

impact tests [19, 20]; however, the validity of this model to reproduce the high-velocity

impact behavior of foam cores has yet to be studied in proven.

The goal of this research is to develop a simple constitutive model that will

describe the behavior of the sandwich core material subjected to cyclic loads. This model

should address the micro-inertia and rate effects, as well as permanent crushing or

damage of cell walls. Chapter II presents a review of existing constitutive models of

foams, especially the rate-dependent ones. Chapter III describes experiments and includes

the description of a test plan, apparatus design, test setup and experimental results. Two

types of tests are done: compression and shear. The tests are strain-controlled, cyclic tests

5
under uniaxial compression and pure shear loadings, with both in-plane and out-off-plane

directions tested under room temperature. Chapter IV discusses two rheological models:

one addresses viscoelasticity, viscoplasticity together with Mullins damage; the other is

based on the theories of damage mechanics. Conclusions are finally made in Chapter V.

6
CHAPTER II

LITERATURE REVIEW

Many articles are written on the response of foam-core composite sandwich

structures subjected to blast loads [21-24]. Because of the difficulties of getting test

results under blast loads, computational simulations have been widely used. Current

models that utilize crushable elastic-plastic constitutive models to simulate the behavior

of the foam core in the sandwich panels or beams may address the material behavior

under quasi-static monotonic loading. However, they fail to simulate hysteresis/energy

dissipation that take place in polymer foams under dynamic loading characterized by

transient and vibration motion.

In this chapter finite element analysis (FEA) packages for simulating crushable

foams in ABAQUS and LS-DYNA are reviewed. Both the validation and limitation of

these two software packages are discussed and compared. The application of crushable

foam plasticity models implemented in ABAQUS will be further analyzed. Before the

numerical review, this chapter starts with revealing some foam core material behaviors

derived from the experiments, which will be helpfully for later discussion.

7
2.1 Experimental Investigation

A typical compression stress-strain curve for foam can be described by three

regions: a linear elastic region, a plateau region and a densification region. As shown in

Figure 2.1, the stress-strain curve indicates that the material behaves linearly elastic with

a slope equal to the Young’s modulus of the material when the strain is very small. As the

load increases, a stress plateau is generated. This roughly constant load is caused by the

collapse of foam cells. Depending on the mechanical properties of the cell walls, the

collapse may be caused by elastic buckling, plastic yielding or brittle crushing [25]. As

the cell walls keep on collapsing, the opposing cell walls meet and touch each other and

this procedure causes the stress to increase very rapidly with strain in the densification

region. Most current researchers are interested in studying the material behavior before

the densification region.

Figure 2.1 A typical foam compression stress-strain curve [25].

8
For applications, such as in the marine industry, it is necessary to study core

materials in sandwich structures subjected to multiaxial loadings. For modeling purposes,

it should be made clear if strengths are different in tension and compression, and whether

the material is isotropic or anisotropic. A study of the foams under multiaxial stress

conditions was done by Gdoutos et al. [26]. They performed uniaxial tensile,

compressive and shear tests along the in-plane and the out-of-plane directions on

Divinycell PVC H100 and H250 foams. Their results indicated that the elastic modulus

under uniaxial tensile and compressive loadings were almost equal for both Divinycell

PVC H100 and H250. The authors also claimed that the low-density Divinycell PVC

H100 displayed nearly isotropic behavior, while the higher density Divinycell PVC H250

was strongly anisotropic with respect to stiffness and strength. The elastic modulus and

strength of Divinycell PVC H250 along the out-of-plane direction were much higher than

along the in-plane directions. The authors explained this was related to the microstructure

of the material where the cells were elongated in the thickness direction. The failure

surfaces (defined as yield) of Divinycell PVC H250 under combined normal and shear

stresses along the in-plane and out-of-plane directions were obtained from biaxial tests. It

was found that the failure envelopes in the principal plane were ellipses displaced along

the first quadrant of a graph with their long axis closer to the through-the-thickness than

the in-plane axis. This phenomenon was predicted well by the Tsai–Wu failure criterion

typically used in composite materials. Daniel et al. [27] studied the failure of PVC H250

as well, and drew the same conclusion that the Tsai-Wu theory could predict anisotropic

material yielding of the foam.

9
2.2 Finite Element Models for Foam

Based on the current experimental studies on foams, two FEA packages have

been widely utilized to simulate foam core behaviors: ABAQUS and LS-DYNA. In

ABAQUS, there are two well-established foam models: the crushable foam model and

the hyperfoam model [11]. The crushable foam model in ABAQUS is based on the

plasticity theory. It could be used to model foam deformation under compression due to

cell wall buckling process. On the other hand, the hyperfoam material model is based on

the hyperelastic theory, but it could not model the strain rate response of foam. The PVC

foam is simulated by the crushable foam material model implemented in ABAQUS.

The foam model in LS-DYNA is similar to the one in ABAQUS. It allows the use

of a hysteretic unloading factor and a shape factor to represent the hysteretic behavior of

an elastomer foam material [28]. For crushable foam, hyperfoam, and low density foam,

the model only accepts test data of one stress-strain curve at a certain strain rate. The

modified crushable foam model in LS-DYNA and Fu-Chang foam model [29] accept test

data for stress-strain curves at differing strain rates. However, Pan [30] claimed the Fu-

Chang rate responsive material foam model could not be captured well under high strain

rate, strain rate above 50 s-1.

To evaluate and compare the validation of the two FEA packages, Ozturka and

Anlasb [31] applied both ABAQUS and LS-DYNA FEA simulations to study multiple

compressive loading and unloading with polystyrene (EPS) foam. They used crushable

foam material model in ABAQUS and low density foam material model (MAT 57) in

10
LS-DYNA. The research reported that both packages were accurate in predicting

maximum deceleration, force, and displacement during the first loading. However, the

authors claimed that the FEA packages need improvement for the case of unloading and

reloading. For instance, they observed that ABAQUS over-predicted maximum

deceleration, force, and displacement for the same amount of absorbed energy at each

loading step. It was also found that LS-DYNA would provide accurate results if

parameters for controlling shape and hysteresis of unloading and reloading were

calibrated using test results. The force and displacement needed to be improved in the

unloading region and so did the residual deformation after unloading.

2.2.1 Quasi-static and low-velocity impact

For decades, researchers struggle to set up reliable FEA models that could

accurately simulate foam behavior under both quasi-static and impact loading conditions.

It is found that the crushable foam model implemented in ABAQUS could perform well

under quasi-static and low-velocity impact before unloading. Schubel et al. [32] studied

composite sandwich structures that subjected to low velocity impact damage. The

experimental setup included sandwich panels consisted of woven carbon/epoxy face-

sheets and a PVC foam core, and instrumented panels impacted with a drop mass. The

comparison between low velocity impact and an equivalent static loading indicated that

regardless of localized damage, low velocity impact was generally quasi-static in nature.

Therefore, a quasi-static test, which is easier to perform and analyze, could be used to

predict related low velocity impact. In the Ref. [32], experimental results were also

compared with the model constructed in ABAQUS. The facesheets were modeled as

orthotropic laminate and the core was modeled as an isotropic elastic material until

11
yielding. After core yielding, the core was modeled as crushable foam. Finite element

analysis (FEA) using an elastic foundation compared well with experiments in the linear

region before the core yielded. However the researchers claimed that there was a need for

a FEA model to capture the response of the panel behavior after yielding.

Rizov [33] also applied the crushable foam and the crushable foam hardening

options of ABAQUS code to the plastic response of the foam core. He developed a 2D

finite element modeling procedure for analyzing the non-linear indentation behavior of

foam-cored sandwich composite beam. The indenter was modeled as a rigid cylinder, and

the facesheets were assumed linear-elastic and quasi-isotropic. The core was modeled as

an elastic-plastic material with isotropic hardening. The elastic-plastic model was

obtained from uniaxial compression test in terms of true (Cauchy) stresses and volumetric

logarithmic plastic strains. The experimental data and the numerical analysis showed very

good agreement with the finite element modeling and could be used to predict the static

indentation response of foam cored sandwich beams. However, during the test it was

observed that after unloading the magnitude of the residual dent decreased with time. The

author claimed that it was due to the relaxation of the crushed foam core. For that reason,

it was concluded that the model was capable of predicting the residual dent at the instant

of impact.

2.2.2 High-velocity impact

High-velocity impact behavior is different from the low-velocity one, and the

response of the foam under high-velocity impact is hard to obtain. According to a

comprehensive review by Abrate [6], high-velocity impact is defined as where the ratio

12
between impact velocity and the velocity of compressive waves propagating through the

thickness is larger than the maximum strain to failure in that direction. Abrate indicated

this implied that damage was generated during the first few travels of the compressive

wave through the thickness when overall plate motion was not yet established. Thus, it

could be concluded that high-velocity impact is a phenomenon controlled by wave

propagation. Consequently the conclusions drawn in studies on static or low-velocity

impacts ware not applicable to high-velocity cases. Since the experiments under high-

velocity impacts are difficult, the theoretical models concerning them are immature and

not well-established.

Ivañez and Santiuste [34] studied the high-velocity impact response of sandwich

plates, with E-glass fiber/polyester face-sheets and foam core. They used 3D finite-

element model in ABAQUS Explicit to simulate the material behavior. For the

facesheets, they implemented Hou failure criteria [35] and a procedure to degrade

material properties in a user-material subroutine (VUMAT). For the foam-core, they

modeled it as a crushable foam material. The author concluded that under certain impact

velocities (280 in/s), the facesheets absorbed most of the impact energy and the foam

core’s influence was negligible. Fiber failure was found to be the main failure mechanism

in the composite facesheets, and it was responsible for the damaged area produced in the

facesheets.

2.2.3 Cyclic loading

Recent technological advances have been applied to increase the crushing strength

of polymer foams. If the facesheets and foam core do not fracture and are still bonded

13
together well, the foam-core will experience load-unload cycles and hysteresis due to

relative motion of facesheets. To better understand how hysteresis develops in the foam,

two separate finite element analyses of glass/vinylester-PVC H100 foam core sandwich

constructions are considered. They are described in Figure 2.2 and Figure 2.3, as a

projectile impact into a sandwich panel and blast loading of a sandwich shell,

respectively. The facesheets are orthotropic elastic and the PVC core is modeled as

crushable foam with isotropic hardening [36].

2
Compressive Stress (MPa)

1.5
1

A● 0.5
B● 0 A
C● -0.5
B
C
-1
-1.5
-2
-0.05 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Compressive Strain

(a) (b)

Figure 2.2 Projectile impact into sandwich panel:


(a) FEA results and (b) Hysteresis in the foam.

14
2

Compressive Stress (MPa)


1.5
1
0.5
0
-0.5 A
-1 B
-1.5
-2
-0.12 -0.08 -0.04 0 0.04 0.08 0.12 0.16
Compressive Strain

(a) (b)

Figure 2.3 Blast loading of a sandwich shell:


(a) FEA model and (b) Hysteresis in the foam.

The compressive stress-strain responses at various locations in the foam are

shown for the impacted sandwich panel and the blast-loaded sandwich shell in Figures

2.2(b) and 2.3(b). Elastic-plastic waves propagate in the foam and reflect off the

interfaces within the sandwich structures. The elastic waves travel faster because of the

initial high modulus of the foam before it yields. There is hysteresis due to plasticity but

no viscoelasticity. The FEA simulations revealed some important facts about the foam.

The propagation of stress waves within the foam is controlled by the dynamics of the

facesheets and structure as a whole. In the case of projectile impact, the foam is

compressed beyond the crushing (yield) strength of the foam in the impact zone (Points A

or B in Figure 2.2 (a)) but this compressive stress is relieved when the projectile

15
rebounds. Tensile residual stresses and compressive permanent strain remain in the

impact region of the foam. Point C in Figure 2.2 (a) experiences only elastic vibrations.

2.3 Constitutive Models for Foam

2.3.1 Crushable foam model

Plastic deformation in a fully dense metal occurs at constant volume. Because of this, the

yield criterion that characterizes the plastic behavior in metals is independent of mean

stress. Deshpande and Fleck [37], who studied the yield criterion and the plastic response

of metal foams under various combinations of axial and radial compression, tension, and

shear loading, found that the yield surface in the stress space is adequately described by a

quadratic function of mean stress versus effective stress, with the hydrostatic yield

strength comparable to the uniaxial yield strength. Deshpande-Fleck claimed their foam

plasticity model could be applied for the open and closed-cell aluminum alloy foams.

They assumed the material is isotropic and the yield criterion is given as

̂ (2.1)

where Y is the yield strength and the equivalent stress ̂ is defined by

̂ [ ] (2.2)
[ ( ) ]

where is the effective stress or Mises stress, is a hydrostatic strength given by

√ ⁄
| | (2.3)

16
and is a parameter that defines the aspect ratio of the ellipse in the - space. When

̂ reduced to . The value of is greater than zero and lies in a certain range.

For instance, in Alporas and Duocel foams, is in the range 1.35 to 2.08. The parameter

could be derived in terms of plastic Poisson ratio as


(2.4)

After defining the shape of the yield surface, the next step is to describe how the

yield surface evolves with strain. For an isotropic material, the applied associated flow

rule, which means that the change of stress with plastic strain, is the same in all

directions. Hence, the plastic strain rate is normal to the yield surface or

̇ ̇̂ (2.5)

where ̇̂ is the equivalent plastic strain rate, and it is the work conjugate of ̂, such that

̂ ̂̇ ̇ (2.6)

From Equations (2.1) and (2.2), one gets

( ⁄ ) [ ̂ ̂
] (2.7)

Therefore, the equivalent plastic strain rate is given explicitly by

̂̇ ( ⁄ ) [ ̇ ̇ ] (2.8)

17
A hardening curve is specified by the evolution of ̂ with ̂ . The uniaxial σ-ε curve is

often used to define ̂- ̂ .The true stress σ and logarithmic (true) plastic strain , may be

defined as

̇ ̇⁄ (2.9)

where the slope h evolves with increasing stress level σ.

The crushable foam plasticity models in ABAQUS are intended for analyzing the

foams that are typically used as energy absorption structures. Two hardening models for

foam plastic behavior are available: the isotropic hardening model and the volumetric

hardening model. Both models use the von-Mises yield surface with an ellipse that

depends on deviatoric stress and meridional stress (p-q). The hardening curve describes

the uniaxial compression yield stress as a function of the corresponding plastic strain, and

the increment of plastic strain could be described as the plastic flow rule ( ̇ ̅̇ ).

The true (Cauchy) stress and logarithmic strain values should be given in defining the

dependence at finite strains.

As shown in Figure 2.4, the yield ellipse of the isotropic hardening model is

centered at the origin in the p–q stress plane. The yield surface evolves in a geometrically

self-similar manner. This phenomenological isotropic model was the one originally

developed for metallic foams by Deshpande and Fleck [9] and described earlier.

18
Figure 2.4 Crushable foam model with isotropic hardening:
yield surface and flow potential in the p–q stress plane [11].

The volumetric hardening model was proposed by Wilde [41] as shown in Figure

2.5, where the point on the yield ellipse in the meridional plane that represents

hydrostatic tension loading is fixed and the evolution of the yield surface is driven by the

volumetric compacting plastic strain.

Figure 2.5 Crushable foam model with volumetric hardening:


yield surface and flow potential in the p–q stress plane [11].

Both models predict similar behavior for compression-dominated loading.

However, under tension loading, the isotropic hardening model predicts the same

19
behavior in compression, while volumetric hardening model assumes a perfectly plastic

behavior, with no evolution of the yield surface.

Current crushable foam plasticity models in ABAQUS could simulate the material

response under monotonic loading very well, but lack of the ability for simulating

hysteresis. Figures 2.6 (a) and (b) are illustrations of the stress-strain responses for

current crushable foam models. For both isotropic hardening and volumetric hardening,

yielding occurs with the initial yield stress of . For isotropic hardening, plastic flow

occurs after yielding up to ; this value affects the yield stress during unloading so that

it is a negative . For volumetric hardening, the yield stress in tension does not change

with plastic strain, but the yield stress for unloading is negative .

20
(a) (b)

Figure 2.6 Stress-strain responses of crushable foam models:


(a) Isotropic hardening and (b) Volumetric hardening.

In practice, current crushable foam plasticity models in ABAQUS cannot simulate

the Bauschinger effect, modulus reduction or modulus degradation. The Bauschinger

effect [42] refers to a property of materials where the material's stress/strain

characteristics change as a result of the microscopic stress distribution of the material.

Figure 2.7 illustrates the Bauschinger effect with isotropic hardening. The result indicates

that the unloading yield stress drops with Bauschinger effect to negative . For

example, an increase in tensile yield strength occurs at the expense of compressive yield

strength.

21
Figure 2.7 Bauschinger effect.

Figures 2.8 and 2.9 present the phenomena of modulus reduction and modulus

degradation, respectively. Figure 2.8 (a) is the modulus reduction shear stress- strain

( curve. It indicates that each time the material is loaded with an increasing shear

strain amplitude, the shear modulus ) decreases. This may be attributed to the changes

in the material's microstructure, for instance, the breakage and recovery of physical bonds

linking adjacent clusters. Figure 2.9 (a) is the shear stress- strain ( curve of one

specimen with a constraint shear strain amplitude. After each cycle, the material is

22
damaged. Therefore, the shear modulus ) decreases during the cyclic loading. The

material has its highest shear modulus after being loaded the first time. Both Figure

2.8 (b) and Figure 2.9 (b) indicate that the ratio of shear modulus ( will decrease

with the increasing amplitude of shear strain and number of cycles.

The unloading slope for both isotropic hardening model and volumetric hardening

are constant values, the Young’s modulus. Therefore, current crushable foam plasticity

models in ABAQUS does not simulate a change in the elastic modulus.

(a)

(b)

Figure 2.8 Modulus reduction:


(a) Shear stress-strain ( and (b) Shear modulus ratio verses shear strain.

23
(a)

(b)

Figure 2.9 Modulus degradation:


(a) Shear stress-strain ( and (b) Shear modulus ratio verses number of cycles.

The theories for metallic foams do not describe the polymer foam accurately.

There are different hystereses in the metallic and the polymer foams after compressive

yielding as shown in Figures 2.10 (a) and (b), respectively. The difference between

metallic foam and polymer foam is the unloading response, which in the metallic foam is

elastic but in the polymer foam is viscoelastic. In polymer foams, viscoelastic energy is

dissipated during intermittent loading-unloading cycles in addition to plastic dissipation

(permanent compressive strain) due to the collapse of cell walls. The PVC foam exhibits

not only viscoelasticity but also a Mullins effect. Such history effects have been well

documented for rubbers but not yet for polymer foams.

24
Compressive Cauchy axial stress (Mpa)
True axial strain -
(a) (b)

Figure 2.10 Effect of strain rate on the compressive behavior:


(a) Idealized metallic foam and (b) Divinycell PVC H200 foam [38].

2.3.2 Viscoelasticity

The viscoelastic approach has been widely applied to model time-dependant

behavior of materials, such as metals, polymers, soils, concrete, rocks, etc. Viscoelasticity

consists of an elastic component (E) and a viscous component (η). As shown in Figs.

2.11 (a), (b) and (c), the Maxwell model, and the Voigt model are two phenomenological-

based models. The linear spring-Maxwell (Figure 2.11 (c)) is a three-parameter model

with a Maxwell element in parallel with an elastic spring. Unlike elastic materials, which

do not dissipate energy or have hysteresis after loads are removed, viscoelastic materials

do. Hysteresis is observed in the stress-strain curve, with the area of the loop being equal

to the energy lost during each loading-unloading cycle. The two main characteristics

associated with viscoelastic materials are stress relaxation and creep. Viscoelastic

25
behaviour may be modeled based on experimental results observed in a creep test or a

relaxation test.

(a) (b) (c)

Figure 2.11 Three viscoelastic models:


(a) Maxwell; (b) Voigt and (c) Linear Spring-Maxwell model.

Holzapfel and Simo [39] developed a thermomechanical model of finite

deformation viscoelasticity which provided significant influence for the later work in this

field. The theory is compatible with the second law of thermodynamics in form of the

Clausius Duhem inequality. However, the nonlinear rate dependence and equilibrium

hysteresis effects were not incorporated in their model.

Viscoplasticity theory actually is an extension of viscoelastic model with a

permanent strain. Ti et al. [40] reviewed various constitutive models that have been

proposed to describe the rheological behavior of clay, which is elasto-viscoplasticly.

Their study indicated that for simple analysis of modeling elasto-viscoplastic material, a

perfectly plastic model could be applied.

26
2.4 Energy Dissipation under Cyclic Loading

The stress-strain relation from preliminary experiments indicated that PVC foam

exhibit similar behavior as rubber under cyclic loadings, which is known as Mullins

effect. The Mullins effect was first investigated and studied by Mullins and his co-

workers back in 1947 [57]. Also known as Mullins softening, the Mullins effect is a

damage effect usually used to describe the mechanical response of rubbers-like material

under cyclic loadings. It results in a lower stress for the same strain after material

reloaded of each cycle. As shown in Figure 2.12, the closed loop of stress-strain curve

represents the hysteresis or the energy lost which is an irreversible phenomenon. The

stress t is plotted against λ, which is the eigenvalue of the Cauchy-Green tensor. The

curve is the primary loading plot under simple tension. Considered the loading

path which is the primary loading terminated at an arbitrary point , then unloading

followed path Ba. When the material is reloaded, the path is retraced as aB , and if

further loading is applied, it followed the path c, which is the continuation of the

primary loading path after . If further loading is applied to then the path Ca is

followed on unloading and then retraced back to on reloading. If no further loading

beyond is applied then the curve aC represents the subsequent material response,

which is then elastic. For loading beyond , the primary path is again followed and the

pattern described is repeated. Mullins effect, together with large deformation, non- linear

and hyperelastic behavior become the complexity of the mechanical behavior of rubbers.

27
Figure 2.12 Schematic loading-unloading curves in simple tension [45].

Researchers have proposed several physical interpretations of the Mullins effect.

These may vary from a micro-mechanics level, such as the broken chain at the interface

between the rubber and the fillers, slipping of molecules, rupture of the clusters of fillers,

chain disentanglements, to more complex composite structure formation. [45]

Most of the studies of Mullins effect are on rubbers under uniaxial tension as

opposed to multi-axial tension. Some remarkable articles in the literature are Ogden and

Roxburgh (1999) for Mullins effect [45] and the extension of this model by Dorfmann

and Ogden (2004) which includes the permanent set phenomenon. [58].

28
In Simo’s method[56], the strain energy function is augmented by a damage

function :

(4.31)

where is the usual description of the strain energy function.

In the Ogden-Roxburgh method[45], damage is given by a damage parameter η

(4.32)

Both and are scalar functions that can be used to define microstructural damage of

the foam, such as microbuckling and even fracture at bending hinges.

Traditional Mullins effect theories do not consider the residual strain upon

unloading as the mechanical properties of rubber were represented in terms of a strain-

energy function. Dorfmann and Ogden [58] proposed a stress softening model in rubber

together with associated residual strain effects. Their results showed how the stress

softening and residual strain change with the magnitude of the applied strain. They

modified the original Ogden-Roxburgh strain energy function by adding a new function

to describe Mullins effect with residual strain. The damage function includes two

variables in the energy function in order separately to capture the stress softening and

residual strain effects:

(4.33)

29
where N is the function introduced to characterize the residual strains and , are the

dissipation functions that are subjected to (1) = 0, (1) = 0. Note that when = 1,

Equation (4.33) becomes analogue to Equation (4.32).

2.5 Combination of the Existing Theories

The need of understanding complex response of material behavior is increasing,

which leads to the investigation of theories that could combine the existing theories for

different physical phenomena. Since individual models were developed within the pre-

defined requirements, coupled formulations could be constructed by simply combining

models. Coupled elastoplasticity with continuum damage mechanics (CDM) has been

investigated back in 1987 by Simo and Ju [46] and later by Yazdani and Schreyer [47].

Hansen and Schreyer [48] developed a rigorous thermodynamic framework with coupled

elastoplastic and damage theories. It was the combination of anisotropic plasticity and

anisotropic damage formulations. The authors claimed that the concept of effective stress

was the critical mechanism for coupling these theories. The coupled formulations were

applied to an aluminum alloy, and it was found that the results were as expected. Borst et

al. [49] investigated two theories: one was the combination of gradient plasticity with

scalar damage and the other was combination of isotropic plasticity with gradient

damage. The article discussed the physical relevance of coupling hardening/softening

plasticity with damage governed by different damage evolution functions. It also

mentioned the sensitivity of the results with respect to the discretization and analyzed

some model parameters. The two theories were valid for finite element simulations of

localization in a one-dimensional model problem, and the model which combined

30
gradient-damage with hardening plasticity could be used to predict fracture mechanisms

in a compact tension test.

31
CHAPTER III

EXPERIMENTS

In order to formulate a three-dimensional constitutive model for polymer foams,

material data are needed in various test modes. Special apparatus for determining the

properties of Divinycell PVC H100 foam in monotonic compression, cyclic compression

and pure shear tests in both out-of-plane and in-plane directions are designed. This

chapter begins with an introduction of Divinycell PVC H100 foam and then describes the

test plan, apparatus setup and results.

3.1 Specimens

The foam used for the tests in this study is Divinycell PVC H100. Table 3.1 lists

the material properties provided by DIAB [2]. Divinycell PVC H100 is a solid closed-

cell, cross-linked foam with a average cell size of 400μm.

32
Table 3.1 Characteristics Divinycell H100 [2].

The PVC H100 material was purchased from the Aircraft Spruce & Specialty

Company in sheets. As shown in Figure 3.1, the though-thickness or

3-direction is defined as the out-of-plane direction, and the orientation of 1- and 2- are

defined as the in-plane directions. The specimens were cut by a slitting saw operated on

Bridgeport mill machine. Examine four specimens cut from the sheet as shown in Figure

3.1. If the load is applied along the 3-direction of the cubic specimen (left lower corner),

the test is an out-of-plane compressive test. When the load is applied along either 1- or 2-

direction of the cubic specimen (right lower corner), the test is defined as an in-plane

compressive test. If the load is applied along the solid arrow directions of the rectangular

specimen (right higher corner), the test is an out-of-plane shear test. When the load is

applied along the open arrow directions of the block specimen (left higher corner), the

test is defined as an in-plane shear test.

33
Figure 3.1 Divinycell PVC H100 material sheet.

3.2 Test Plan

Monotonic compressive tests in the out-of-plane direction were done first to

verify the manufacturer-published properties [2] and analyze the strain effect on PVC

H100 foam. In this research, the strain is limited to 20%, long before the densification

region, and strain rates from 10-4-0.1 s-1. At a strain rate of 0.1 s-1 , the monotonic

compressive test in the in-plane directions is compared with that of the out-of-plane

direction to show that the material is transversely isotropic.

Cyclic compression tests and pure shear tests in both the out-of-plane and in-plane

directions were also done under different strain amplitudes and strain rates. The strain

rates were varied from 5×10-4 to 5 s-1, over five orders of magnitude. For each strain rate

of cyclic compression tests, the strain amplitudes ranged from 2-10% with 2%

increments. For the pure shear tests, the strain amplitudes ranged from 4-20% with 4%

34
increments on the foam specimen. A parametric study on the combination of strain

amplitude and strain rates that can be achieved by adjusting the PID (Proportional–

Integral–Derivative) controller of MTS 831 servo-hydraulic machine was done to satisfy

the test plan.

The cyclic compression strain history at strain rate equal to 5 s-1 with a strain

amplitude equal to 4% is shown as shown in Figure 3.2. The strain rate is given by

, where and The test frequency is . Therefore, at strain

rate equal to 5 s-1 and strain amplitude equal to 4%, the MTS machine was required to

operate at a frequency of 62.5 Hz. For the same strain rate, the 2% strain amplitude test

would require the MTS machine to operate at a frequency of 125 Hz, etc. In addition to

the constant strain rate requirement, the waveform has to be a sawtooth shape and it is

very difficult to tune the MTS machine with this type of wave form at high frequency.

The PID tuning parameters of the MTS machine were adjusted of produce the correct

waveform over a range of strain amplitudes and frequencies. Finally, to collect enough

data points for the output files and capture the correct waveform at high frequency, the

data acquisition was also adjusted. Take again as an example, a strain rate equal to 5 s-1

with a strain amplitude equal to 4%, as demonstrated in Figure 3.2. If one requires 100

points during the first loading cycle, the data acquisition should be set at or ,

which means the time between two points is 0.00008 s or data per second.

35
T
ε
∆t
t

𝜀 5
-0.04

Figure 3.2 Strain history at a strain rate 5 s-1 and strain amplitude 4%.

Table 3.2 and Table 3.3 show the test frequencies for the cyclic compression test

and pure shear test, respectively, after tuning the MTS servo-hydraulic machine.

Table 3.2 Test frequency of MTS machine in cyclic compression tests.

Strain Rate (s-1) Strain Amplitude (%)

2 4 6 8 10

0.0005 0.0125 Hz 0.00625 Hz 0.00417 Hz 0.00313 Hz 0.0025 Hz

0.0050 0.125 Hz 0.0625 Hz 0.0417 Hz 0.0313 Hz 0.025 Hz

0.0500 1.25 Hz 0.625 Hz 0.417 Hz 0.313 Hz 0.25 Hz

0.5000 12.5 Hz* 6.25 Hz 4.17 Hz 3.13 Hz 2.5 Hz

5.0000 125.0 Hz* 62.5 Hz* 41.7 Hz 31.3 Hz 25.0Hz

*Cyclic test under these combinations of strain amplitudes and frequencies could not
be achieved by the controller.

36
Table 3.3 Test frequency of MTS machine in pure shear tests.

Strain Rate (s-1) Strain Amplitude (%)

4 8 12 16 20

0.0005 0.00625 Hz 0.00313 Hz 0.00208 Hz 0.00156 Hz 0.00125 Hz

0.0050 0.0625 Hz 0.0313 Hz 0.0208 Hz 0.0156 Hz 0.0125 Hz

0.0500 0.625 Hz 0.313 Hz 0.208 Hz 0.156 Hz 0.125 Hz

0.5000 6.25 Hz 3.13 Hz 2.08 Hz 1.56 Hz 1.25 Hz

5.0000 62.5 Hz 31.3 Hz 20.8 Hz 15.6 Hz 12.5 Hz

3.3 Compressive Tests

3.3.1 Apparatus design and test setup

The apparatus and setup for monotonic tests and compression cyclic tests are

basically the same. For monotonic tension tests and cyclic compression tests, the

specimen needs to be glued onto aluminum grips as shown in Figure 3.3 because of the

tensile loads. The adhesion agent used was LOCTITE® Quick SetTM. It can provide a

maximum of 3,211 psi shear strength on sand blasted, cold rolled steel after 24 hours of

curing [59]. It comes in two separate bottles. One is the resin, and the other is the

hardener. Every time the solvents should be mixed with a 1:1 ratio and blended well with

wood sticks. Experience indicated that the mixture should be used within 1 minute after it

is exposed to air. However, the foremost thing that should be seriously considered is that

the surfaces of foam and aluminum plates should be clean, dry and free of grease or oil.

First one side of the specimen was glued and cured for 30 minutes until the glue set. Then

37
to eliminate any MTS machine misalignment, the other side of the specimen was glued

and pre-set to cure in the MTS machine. After another 30 minutes, the specimen was

taken out of the MTS machine and cured for 24 hours in a room temperature

environment. After each test, the specimen is broken in two parts by MTS machine under

displacement control. The majority of the foam that left on the aluminum grips can be

removed by scraping with blades. Sandpapers are then used to further remove any

residual foam on the grips. Finally, the grip surface is cleaned with alcohol before the

next specimen is attached.

There are two standard test methods for compression/tension tests of foams: one

is ASTM 1621 (Compressive Properties of Rigid Cellular Plastics) and the other is ISO

844 (Rigid Cellular Plastics Compression Properties). In this research, the ISO 844 (Rigid

Cellular Plastics-Determination of Compression Properties) was applied and modified, as

shown in Figure 3.4. The specimen is a 1in x 1in x1in cube, and it was glued onto

aluminum grips, with diameter of 1.5in with 0.75in thickness. The directions of the

specimen are shown in Figure 3.1. A pair of “C” clamp was specially designed to connect

the actuator arm of MTS and the test piece, as shown in Figures 3.5. (a) and (b). The

hollow inside of the “C” clamps is a frisbee shape which can easily hold the actuator arm

of the MTS machine and the test piece. To eliminate the possibility of generating

additional stress along the vertical direction, the “C” clamps were connected by screws

orientated horizontally.

38
Figure 3.3 PVC H100 glued on Aluminum gripes.

Figure 3.4 Set up for compressive cyclic test for PVC H100.

39
(a)

(b)

Figure 3.5 Set up of “C” clamps:


(a) A pair of “C” clamps and (b) Side views of “C” clamp.

3.3.2 Monotonic tests results

The ISO 844 test standard gives a method to define the linear portion in the elastic

region on the stress-strain curve. As shown in Figure 3.6, the nonlinear portions at the

beginning and the end of elastic region are cut off, and the remaining portion is

40
defined as the linear elastic region. The slope of portion is used to define the Young’s

modulus of the material.

Figure 3.6 Procedure to define linear portion in the elastic region.

Figure 3.7 is the stress-strain curve obtained from the test at a strain rate of 0.1 s-1.

The test data was processed following the ISO 844 method mentioned above. As shown

in Figure 3.10, the original test data is the solid black curve, which has nonlinear portions

at the beginning and the end in the elastic region. After the cut off the nonlinear portions,

the Young’s modulus has been defined as 8,940.4 psi. The curve is then shifted to the

origin to get rid of the nonlinear portion at the beginning of the test. The yield strain

associated with the shifted curve is found where the linear portion ends, and this is

presented as the vertical solid red line in Figure 3.7.

Tests were done at various strain rates to show that the material is viscoelastic and

viscoplastic, in Figure 3.8. Here, the viscoelasticity is signified by a slight change in the

41
Young’s modulus as the strain rate increases. The viscoplasticity is signified by the

change in the yield strength or flow stress as the strain rate increase.

Figure 3.7 Out-of-plane monotonic test under strain rate of 0.1 s-1.

Figure 3.8 Out-of-plane monotonic test at various strain rates.

42
Table 3.4 summarizes the strain rate effect for the material. The actual stress-

strain curves at different strain rates are given in Appendix A. As strain rates increase,

compressive modulus (slope of the elastic portion), compressive strength (stress at the

yield point) and plateau stress (stress in the plateau region) increase. For different strain

rates, yielding occurred at around 2% strain.

Table 3.4 Out-of-plane strain rate effect for PVC H100.

US Unit
Strain Rate Compressive Modulus Compressive Strength Plateau Stress Yield Strain
(s-1) (psi) (psi) (psi) (%)
0.0001 7629.9 237 240 3.06
0.0010 8506.1 259 256 3.00
0.0100 8780.9 268 272 3.02
0.1000 8940.0 289 295 3.08

With respect to the material directions defined in Figure 3.1, tests indicated the

material behaves the same in the 1- and 2- directions, but the stress-strain curve in the 1-

and 2-directions are different from that in the 3-direction. Hence, it could be concluded

that PVC H100 is a transversely isotropic material. Figure 3.9 shows a comparison of the

stress-strain curves in the out-of-plane and in-plane directions. It is obvious that the

modulus and strength are different. The strength in the in-plane direction is 40% lower

compare to the out-of-plane direction.

43
Figure 3.9 Out-of-plane vs. in-plane at strain rate of 0.1 s-1.

Comparison of glued and unglued specimens

Experimentalists question about the effects of gluing the foam specimen on

aluminum grips. Figure 3.10 shows the difference between these two situations at strain

rate of 0.1 s-1. The compressive modulus of the glued test (13,702 psi) is higher compared

with the unglued test (8,940 psi). The compressive strength (stress at yield), as shown by

Points A and B in Figure 3.10 occurred at the same stress value. Plateau stresses are

different, but within 5% of each other.

44
A
B

Figure 3.10 Out-of-plane compressive stress-strain curves


for glued and unglued specimens at strain rate of 0.1s-1.

To verify the accuracy of test setup, the results has been compared to material

properties provided from DIAB in Table 3.1. Notice that DIAB’s tests were based on

ASTM standard D1621, which is different from the ISO 844 that was used in this

research. Klegecell R grade-Performance Characteristics sheet [50] indicated that

compressive modulus will be doubled following ASTM D 1621 when comparing to ISO

844. The ASTM D 1621 is the test standard under strain rate of 0.00167 s-1, and the

results obtained from the test standard in Table 3.1 are for unglued specimens under

different strain rates. An assumption was made to obtain the compressive modulus for

glued specimen at a strain rate of 0.001 s-1. Assume a scalar exists for compressive

modulus for glued and unglued conditions. The scalar is assumed to be the ratio of

compressive modulus for glued over compressive modulus for unglued. As displayed in

Figure 3.10, the scalar k=13,702/8,940. Apply this scalar to strain rate of 0.001 s-1, the

45
glued compressive modulus could be calculate out as 8,506 psi (Table 3.4), which is

roughly equals to 13,000 psi. Double this value and compare it with the data listed by in

Table 3.1. There will be a 30% over prediction for glued specimen at strain rate of 0.001

s-1. This difference is due to the size effect of the test specimen.

The tensile modulus has been compared with the compressive modulus after the

foam glued on the aluminum grips. As shown in Figure 3.11, tensile modulus is 15%

higher than compressive modulus. In the later study, this difference will be neglect.

Figure 3.11 Out-of-plane tensile test vs. compressive test.

3.3.3 Cyclic compression tests results

As mentioned earlier, the PVC H100 was glued to aluminum grips so that the

specimen could be loaded in compression and tension. The cyclic compression tests in

both out-of-plane and in-plane directions were done on PVC H100 foams and the stress-

46
strain curves of different strain amplitudes at various strain rates were obtained. Details

of the results from these tests are provided in Appendix B. The stress-strain response at

strain amplitude of 10% at different strain rates in out-of-plane and in-plane are shown in

Figure 3.12 and Figure 3.13, respectively. The experiments indicated that the PVC H100

exhibits not only viscoelasticity but viscoplasticity with a permanent (plastic) strain upon

unloading to a zero stress state in both the out-of-plane and in-plane directions.

Figure 3.12 Out-of-plane compressive stress-strain curves


at strain amplitude of 0.10 with various strain rates.

47
Figure 3.13 In-plane compressive stress-strain curves
at strain amplitude of 0.10 with various strain rates.

Figure 3.14 is the stress-strain curves of different cycles for strain rate of 0.5 s-1,

with strain amplitude of 0.1 in the out-of-plane direction. Figure 3.15 is the stress trend

for strain rate of 0.5 s-1 with various strain amplitudes in the out-of-plane direction. Both

figures indicated that PVC H100 has modulus degradation, as reviewed in Chapter II.

Figure 3.14 Out-of-plane compressive stress-strain curves for various cycles.

48
Figure 3.15 Out-of-plane compressive stress-strain curves
at strain rate of 0.5 s-1 with various strain amplitudes.

The cyclic behavior with strain rates and strain amplitudes in both out-of-plane

and in-plane directions are shown in Figure 3.16 (a) and (b). Ten loading-unloading

cycles were applied at strain amplitude of 2% with strain rate of 0.5 s-1, and this was

repeated several times with increasing compressive strain amplitudes to a maximum of

10%. In both out-of-plane and in-plane directions, it is noted that the first cycle of the

stresses at the same strain amplitude is higher than those following cycles, which means

damage to the foam occurred after yielding (history effects).

Figure 3.17 shows the first cycle in out-of-plane direction at a strain rate of 0.5s-1

with various strain amplitudes, superimposed on the data with consecutively increasing

strain amplitudes. After the first reloading at strain amplitude of 2%, the stress-strain

response went back to the primary curve beyond 2%. This phenomenon repeats itself for

the different strain amplitudes. The phenomenon shown in Figure 3.17 is described as

Mullins damage [11]. The hysteresis curves of the fourth cycle in the out-of-plane

49
direction in these tests are given in Figure 3.18 to indicate rate-dependent behavior even

after Mullins damage. Similar responses were also found in the in-plane cyclic

compressive tests.

(a)

(b)
Figure 3.16 Compressive stress-strain curve at a strain rate of 0.5 s-1
with various strain amplitudes: (a) out-of-plane and (b) in-plane.

50
Figure 3.17 Out-of-plane first cyclic of strain rates of 0.5 s-1
with various strain amplitudes over lamping on the primary curve.

Figure 3.18 Out-of-plane fourth cycle hysteresis at 10% strain at various strain rates.

51
3.3.4 Material homogeneity

In obtaining mechanical properties, the specimen must be in a homogeneous

deformation and stress state. The ASTM and ISO standards mentioned earlier achieve

this requirement. However, these standards were written for monotonic, quasi-static

response. In this study, we required cyclic, rate-dependent response. Therefore, the

question arises: do the specimens in the experiments deform in a homogeneous manner?

The PVC H100 cubic specimens were loaded in both the in-plane and out-of-

plane directions with strain amplitude of 10% at the lowest (0.0005 s-1) and highest (5 s-1)

strain rates, and the results are shown in Figures 3.19 - 3.22. With a magnification factor

8, specific regions at the top (A1), center (A2), bottom (A3) and the left/right midsection

(A4/A5) were examined for the damaged micro-structures. A comparison of the

magnified results from Figure 3.19 indicated that at strain rate of 0.0005 s-1, the damaged

cells are uniform in the in-plane direction with strain amplitude of 10%. A similar

conclusion is made for the out-of-plane direction shown in Figure 3.20. At strain rate 5 s-
1
and stain amplitude of 10%, the damaged microstructure is also uniform throughout the

specimen in both the in-plane and out-of-plane directions, shown in Figure 3.21 and

Figure 3.22, respectively. Therefore, the conclusion is that PVC H100 deformed

uniformly at strain amplitude of 10% with strain rates from 0.0005 s-1 and 5 s-1 in both

the in-plan and out-of-plane directions.

The permanently crushed specimens when loaded to 10% strain amplitude in the

in-plane direction at five increasing strain rates are shown in Figure 3.23. Here it can be

52
seen that more permanent strain results at higher loading rates. A similar trend appears

for the out-of-plane direction, as shown in Figure 3.23. The damage to the specimen

appears to be history-dependent.

For higher strain amplitudes, the specimen did not crush in a homogenous

manner. It was observed that when the strain amplitude increased to 30% at strain rate of

0.0005 s-1, the material was not crushed homogeneously. The crushed microstructure at

30% strain in the in-plane and out-of-plane directions are shown in Figures 3.25 and 3.26.

Both the top (A1) and bottom (A3) cells were crushed more than at the mid-center (A2)

and sides (A4/A5). This non-uniformity was also observed at larger strain rate of 5 s-1 ,

although results are not show here.

53
(a)

(b)

(c)

Figure 3.19 PVC H100 crushed with 10% strain amplitude at strain rate of 0.0005 s-1 in
the in-plane direction: (a) Crushed specimen (b) Undamaged micro-structure at A2 and
(c) Damaged micro-structures at various locations.

54
(a)

(b)

(c)

Figure 3.20 PVC H100 crushed with 10% strain amplitude at strain rate of 0.0005 s-1 in
the out-of-plane direction: (a) Crushed specimen (b) Undamaged micro-structure at A2
and (c) Damaged micro-structures at various locations.

55
(a)

(b)

(c)

Figure 3.21 PVC H100 crushed with 10% strain amplitude at strain rate of 5 s-1 in the in-
plane direction: (a) Crushed specimen (b) Undamaged micro-structure at A2 and (c)
Damaged micro-structures at various locations.

56
(a)

(b)

(c)

Figure 3.22 PVC H100 crushed with 10% strain amplitude at strain rate of 5 s-1 in the
out-of-plane direction: (a) Crushed specimen (b) Undamaged micro-structure at A2 and
(c) Damaged micro-structures at various locations.

57
0.0005 s-1 0.005 s-1 0.05 s-1 0.5 s-1 5 s-1

Figure 3.23 Crushed specimens loaded to 10% strain amplitude in the in-plane direction
with five increasing strain rates.

0.0005 s-1 0.005 s-1 0.05 s-1 0.5 s-1 5 s-1

Figure 3.24 Crushed specimens loaded to 10% strain amplitude in the out-of-plane
direction with five increasing strain rates.

58
(a)

(b)

(c)

Figure 3.25 PVC H100 crushed with 30% strain amplitude at strain rate of 0.0005 s-1 in
the in-plane direction: (a) Crushed specimen (b) Undamaged micro-structure at A2 and
(c) Damaged micro-structures at various locations.

59
(a)

(b)

(c)

Figure 3.26 PVC H100 crushed with 30% strain amplitude at strain rate of 0.0005 s-1 in
the out-of-plane direction: (a) Crushed specimen (b) Undamaged micro-structure at A2
and (c) Damaged micro-structures at various locations.

60
3.4 Pure shear tests

3.4.1 Apparatus design and tests setup

The design of the pure shear test was based on the ASTM C273 [60] and ISO

1922 [61]. The specimen is a 6in x1in x0.5in block, which is half of the size of standard

specimen. The directions of pure shear tests are defined as shown in Figure 3.1. For each

direction, the 6in x1in surface is glued to the steel plates, which are 6in x 2in x 0.75in

steel block. Unlike the compressive cyclic tests, the specimens in pure shear tests do not

have to be cured on the MTS machine. Because the glued surface area is larger than the

compressive specimen, the curing time of the specimen for pure shear tests was extended

to 48 hours.

The above-mentioned standard test methods are limited to monotonic loading but

not cyclic loading. A specimen test fixture was designed to enable cyclic loading as

showed in Figure 3.27. The design of a 1 in-diameter brass ball encased in aluminum

seats and connected to a stainless steel 304 rod (3/8in-24 threaded) provides freedom of

rotation of the top and bottom edge of the specimen, while eliminating the undesired

displacement. With a safety factor of 2, ABAQUS/Explicit was used to check the failure

of the ball and housing at critical contact regions, such as Point A and Point B as shown

in Figure 3.27 (a).

61
MTS Actuator

Brass Ball

Seats

Steel Rods

Steel Plates

Specimens

Load Cell

(a) (b)

Figure 3.27 Pure shear test apparatus: (a) Schematic of shear text fixture and
(b)Photograph showing experimental setup.

62
Monotonic and cyclic shear tests were performed in the out-of-plane and in-plane

directions as explained in Section 3.2. After each test, the specimen is broken into two

parts while still attached to the plates. Most of the foam will be removed by a cutting

saw. The remaining foam on the plates will then be baked at high temperature. After the

plates are cooled down to room temperature, the burnt foam could simply be scraped off.

Bead blasting was used to take any remaining foam off the steel surface. This is the

process of removing surface deposits by applying fine glass/sand beads at a high pressure

without damaging the surface. Acetone was then applied to do the final cleaning.

3.4.2 Pure shear test results

To verify the accuracy of the results from pure shear test apparatus, current results

were compared to those obtained from Steeves and Fleck [37],the DIAB data sheet [2],

and Daniel [27], respectively.

Because of the similar dimensions of the specimen and test mode, the current test

results were compared with those of Steeves and Fleck (2004) at a strain rate of 0.001 s-1.

As shown in Figure 3.28, the current test result of the out-of-plane pure shear test has the

same plastic flow stress as that found by Steeves and Fleck. However, the two curves

have different shear modulus and different shear fracture strain at load drop. Steeves and

Fleck indicated the shear modulus of PVC H100 was 44 MPa, which is twice of current

out-of-plane shear modulus test of 22 MPa. The shear stress-strain curve in Steeves and

Fleck dropped at 20%, while the current result dropped at 40%.

63
Data from DIAB was used to further validate the current out-of-plane shear test

results. The out-of-plane shear fracture strain was quoted as 40% in the DIAB data sheet,

which is the same fracture strain as found in the current test. However, the shear modulus

of PVC H100 in the out-of-plane direction from the DIAB data sheet was 35 MPa, which

does not match neither that of the current test nor that of Steeves and Fleck (2004).

The current test results were also compared with out-of-plane shear stress-strain

curve from Daniel (2009). Daniel used an Arcan loading fixture and specimen holder to

test PVC H100 under pure shear. As shown in Figure 3.28, the out-of-plane shear stress-

strain curve from Daniel indicated that the shear modulus of PVC H100 was 25 MPa,

which is very close to that in the current test. However, there is no discernible load drop

to indicate a shear fracture strain in the Arcan shear tress-strain curve. This is probably

due to the unique shear stress state imposed by the notched specimen in an Arcan test.

The current test results matched the flow stress of Steeves and Fleck (2004), the

shear fracture strain from DIAB and shear modulus by Daniel (2009). Therefore, the

conclusion that may be drawn here is that the present test results are comparable to those

obtained by independent studies on PVC H100. It was realized that the fracture strain

from Steeves and Fleck (2004) was half of that specified in the DIAB data sheet.

Assuming the DIAB data sheet is a more reliable source, the stress-strain curve from

Steeves and Fleck (2004) was modified by doubling the shear strain on the horizontal

scale. Now the modified stress-strain curve of Steeves and Fleck has the same shear

modulus, the same shear fracture strain and comparable flow stress as the current test

64
result. The difference of current test result versus Steeves and Fleck (2004) could be due

to the definition of the engineering shear strain by Steeves and Fleck. By the test standard

of ASTM C273, the engineering shear strain is the displacement between load plates

derived by the thickness of the core (γ=u/t). It is possible that Steeves and Fleck mistook

the engineering shear strain with tensoral shear strain, the later quantity being half of the

former.

Steeves and Fleck (2004)


Steeves and Fleck
Current test (corrected for engineering
Shear stress, τ13 (MPa)

shear strain)
Daniel (2009)

Shear strain, γ13 (%)

Figure 3.28 Comparison of out-of-plane shear stress-strain curves of PVC H100 foam
core at strain rate =0.001 s-1 with data from Steeves and Fleck (2004) and Danel (2009).

The through-thickness shear fracture of PVC H100 foam core at shear strain rate

of 0.001 s-1 is shown in Figure 3.29. The crack occurred at the center of the specimen and

this indicates that the specimen broke in a state of pure shear and that there was good

bonding between the foam and steel plates. Failure in these tests could be mistaken as

debonding between the foam and steel grips. Clearly this was not the case in these

experiments.

65
Shear Crack

Figure 3.29. Out-of-plane shear fracture of PVC H100 foam core at strain rate =0.001 s-1.

Figure 3.30 shows the out-of-plane and in-plane of monotonic shear stress-strain

response at strain rate=0.5 s-1 with strain amplitude=0.20. The material is again

transversely isotropic. The ratio of the yield stresses of the out-of-plane compare to the

in-plane is approximately 3:2, which is a similar ratio as in the compressive tests, as

shown in Figure 3.9.

66
Figure 3.30 Monotonic shear test at strain rate=0.5 s-1, strain amplitude=0.20.

This stress-strain curve can also be described by three phases: elastic, plastic and

hysteretic with Mullins damage. However, hardening occurs during the plastic flow phase

where the plastic flow stress is not constant.

Cyclic shear stress-strain responses are given in Figures 3.31 and 3.32.

The results of cyclic pure shear tests are similar to cyclic compression test results. As

shown in Figures 3.31 (a) and (b), the cyclic pure shear test in both the out-of-plane

directions and in-plane direction indicate that the material is viscoelastic, viscoplastic and

experience Mullins damage. The pure shear stress-strain curves at one particular strain

rate of 0.5 s-1 with various strain amplitudes in the out-of-plane and in-plane directions

are shown as in Figures 3.32 (a) and (b). More results at different stain rates and various

strain amplitudes are given in Appendix C.

67
(a)

(b)

Figure 3.31 Pure shear stress-strain curves of strain amplitude of 20% at various strain
rates: (a) out-of-plane and (b) in-plane.

68
(a)

(b)

Figure 3.32 Pure shear stress-strain curves at a strain rate of 0.5 s-1 with various strain
amplitudes: (a) out-of-plane and (b) in-plane.

69
3.5 Summary and Conclusions

The main results of current study are summarized as follows:

(1) Monotonic compression stress-strain curves in both out-of-plane and in-plane

directions show that PVC H100 is linear elastic-plastic. The modulus of elasticity

in tension and compression are almost equal.

(2) The PVC H100 foam displays transversely isotropic behavior with the

compressive modulus of elasticity and strength in the out-of-plane direction being

higher than in the in-plane direction. The ratio of out-of-plane to in-plane yield

stress is 3:2.

(3) The modulus and yield strength in monotonic compression stress-strain curves

vary with strain rate and this indicates that the material is viscoelastic and

viscoplastic.

(4) Cyclic compression stress-strain curves derived at different strain rates reveal that

PVC H100 is viscoelastic, viscoplastic and hysteretic with Mullins damage.

(5) Monotonic shear stress-strain curves also indicated that PVC H100 foam displays

transversely isotropic behavior. The same ratio of out-of-plane to in-plane shear

yield stress, 3:2, occurs.

(6) The cyclic shear test results have a similar stress-strain response as the cyclic

compressive tests. The material under cyclic shear is viscoelastic, viscoplastic and

hysteretic with Mullins damage. However, shear hardening occurs after yielding.

70
CHAPTER IV

CONSTITUTIVE MODELING

To simulate the behavior of polymer foam under cyclic compression, two one-

dimensional constitutive models are developed in this chapter. The first model is based

on the combination of viscoelasticity, viscoplasticity and Mullins damage. The second

model is based on the theory of damage mechanics. Only the uniaxial compression mode

is considered in this chapter. Mullins and viscoelastic damage predictions will only be

made for the out-of-plane compressive mode since the in-plane compression mode is very

similar.

4.1 Damage with Mullins Effect

The stress-strain relation of PVC H100 under uniaxial loading indicates that the

material behavior is a combination of viscoelasticity, viscoplasticity and Mullins damage,

as described in Figure 4.1. Based on this material behavior, the phenomenological one-

dimensional constitutive model shown in Figure 4.2 is proposed. Equations for

viscoelasticity, viscoplasticity and Mullins damage are discussed in this section.

The mechanical analogs model is shown in Figure 4.2. An equilibrium spring with

modulus is in parallel to a Maxwell model (spring with modulus and dashpot with

71
η) to simulate viscoelastic response. These are then in series with a Prandtl element with

μ, for viscoplasticity.

Viscoplasticity

Viscoelasticity

Energy Dissipation

Figure 4.1 Material behavior of strain rate of 0.5 s-1 (out-of-plane).

Figure 4.2 Mechanical analogs model.

72
The total stress and strain at any time are given by

(4.1)

and

(4.2)

4.1.1 Viscoelasticity

Figures 4.3 (a) and (b) are the experimental results in both the out-of-plane and

in-plane directions. For each strain rate, the stress-strain relation is linear. The stress-

strain curves also show rate dependency. As strain rate increases, the modulus increases.

This phenomenon indicates the material is viscoelastic.

73
(a)

(b)

Figure 4.3. Experimental stress-strain curves showing viscoelasticity:


(a) out-of-plane and (b) in-plane.

Viscoelastic, as the name implies, is a combination of both viscous behavior

(fluid) and elastic behavior (solid). The simplest mechanical analogs for a viscoelastic

material are Maxwell model and Voigt model (also known as the Kelvin-Voigt model),

which were reviewed in Chapter II. The mechanical analog for viscoelastic in this

74
research is as shown in Figure 4.4. It could be described as a linear spring in parallel with

a Maxwell element.

Figure 4.4 Mechanical analogs for a viscoelastic.

The stress-strain behavior is separated into an elastic component and a rate

dependent viscoelastic component. The elastic component consists of a linear spring with

spring constant of ; the rate dependent viscoelastic component consists another linear

spring with spring constant of , and a nonlinear viscous dashpot. As shown in Figure

4.4, it is clear from the geometry of the model that the total strain will be the strain of the

spring ( plus the strain of the dashpot ( :

(4.3)

The total stress ( is the sum of equilibrium stress ( and overstress ( .

(4.4)

75
The equilibrium stress is described by a spring constant of . The spring with

spring constant of corresponds to the behavior of foam in the most relaxed state,

obtained from lowest strain rate test (strain rate=0.0005 s-1).

(4.5)

The overstress is the change in stress from the equilibrium stress associated with

higher strain rate. It is described by a spring with spring constant

̇ (4.6)

where, is a viscosity coefficient, and which is assumed to be constant in this research.

Substituting Equation (4.3) into Equation (4.6) gives

̇ (4.7)

where . Here, is a constant strain rate, as shown in Figure 4.5. Equation (4.7)

can be rewritten as

- (4.8)

76
ε
2εmax

loadin unloadin reloadin


g g g
εmax

α ε= 2εmax-αt
α
ε= αt ε= -
0 2εmax+αt
t

-2εmax

Figure 4.5 Relationship of strain verses time for loading, unloading, and reloading.

Solving Equation (4.8) with the condition =0, when ε=0, one gets

[ ( ) ] (4.9)

Substituting Equation (4.9) into Equation (4.6) gives

77
[ ( )] (4.10)

Hence, the total stress could be expressed as

[ ( )] (4.11)

As discussed above, is the Young’s Modulus of strain rate of 0.0005 s-1

(lowest strain rate in this research). In the out-of-plane experiments, was found to be

13940 psi, and in the in-plane experiments, was 7916 psi. The value of and η are

the two material properties which need to be defined from the monotonic compressive

stress-strain curves at various strain rates. A nonlinear regression software program,

OriginLab [51], was utilized for curve fitting. Table 4.1 is a summary of the values of

and η for different strain rates in the out-of-plane and the in-plane directions.

Appendix C shows the results from curve fitting.

Table 4.1 Material properties of out-of-plane and in-plane after curve fitting.

Out-of-plane In-plane
Strain rate (s-1) (psi) η (psi·s) (psi) η (psi·s)
0.0005 1980 428 473 56
0.0050 2250 407 486 66
0.0500 2310 539 580 68
0.5000 2430 550 811 78
5.0000 2530 576 650 82
Average 2300 500 600 70

78
Note that there are variable sets of solutions for this nonlinear curve fit. In this

research, the solutions were set in the certain range and values of and η are assumed

to be constant in this research. After some preliminary studies, the average values are

chosen as the final values for and η, as shown in Table 4.2.

Table 4.2 Constant material properties of out-of-plane and in-plane.

Out-of-plane In-plane
(psi) 13940 7916
(psi) 2300 600
η (psi·s) 500 70

Figures 4.6 (a) and (b) are the results of fitted curves. Both the out-of-plane and

in-plane show the trend of viscoelsticity. More details are given in Appendix D. Table 4.3

provides the comparison results of test data to the fitted curves in out-of-plane and in-

plane directions. The comparison errors are within 2%, which is very good.

79
(a)

(b)

Figure 4.6 Fitted curve for viscoelastic response:


(a) Out-of-plane and (b) In-plane.

80
Table 4.3 Comparison of results for viscoelasticity in out-of-plane and in-plane
directions.
Out-of-plane In-plane
Strain rate Peak stress Predicted peak Diff. Predicted stress Predicted peak Diff.
(s-1) (psi) (psi) (%) (psi) (psi) (%)
0.0005 160.29 160.52 0.14 80.26 81.45 1.48
0.0050 162.53 162.2 -0.20 82.68 82.13 -0.66
0.0500 180.89 177.28 -2.00 84.25 84.15 -0.12
0.5000 186.73 186.48 -0.13 85.92 87.33 1.64
5.0000 192.04 188.76 -1.71 88.17 88.01 -0.19

4.1.2 Viscoplasticity

Experiments on metals show that under uniaxial loading, the strain at a given

stress has two parts: a small recoverable elastic strain ( ), and an irreversible plastic

strain ( ), as shown in Figure 4.7.

Figure 4.7 Elastic-plastic stress-strain curve in metals.

Allan [52] reviewed five key concepts in modeling metal plasticity, as below:

1. The decomposition of strain into elastic and plastic parts, .

81
2. Yield criteria, which predict whether the solid responds elastically or

plastically. For instance, for an isotropic metal, yield occurs when von Mises effective

stress ( ) attains the yield value Y, the yield criterion is written as .

3. Strain hardening rules, which control the way in which resistance to plastic

flow increases with plastic straining. There are two simple approaches, one is isotropic

hardening, and the other is kinematic hardening, as shown in Figure 4.8 (a) and (b).

Figure 4.7 (a) indicated that the isotropic hardening law is not useful where components

are subjected to cyclic loading. On the other hand, kinematic hardening law allows the

yield surface to translate, without changing its shape. This type of behavior better

describes the experimental results in this research.

(a) (b)

Figure 4.8 Plastic hardening: (a) Isotropic hardening and (b) Kinematic hardening.

4. The plastic flow rule, which determines the relationship between stress and

plastic strain under multi-axial loading. It is usually expressed as ̇ ̂̇ .

5. The final issue to consider is the elastic unloading criterion. Experiments

show that plastic flow is irreversible, and the material always dissipates energy.

82
The plastic response of foam differs fundamentally from metals. The models for

foams generally differ in three respects (i) the yield criterion; (ii) the strain hardening

law; (iii) the expression of ̂̇ in the plastic flow rule.

A one-dimensional viscoplastic constitutive model [53], which is based on the

theory on rate-independent plasticity under uniaxial loading is applied in this research.

The yield function defined the elastic domain, where the material behavior is assumed to

be purely elastic. The yield function is given by

| | (4.12)

where is the yield stress. When | | , the material behavior is purely elastic. The

strain hardening law describe the changes in yield stress that result from plastic strain. It

could be describe as

̅ (4.13)

In this research, the yield stress was taken to be constant (perfectly plastic

behavior). This is because the compressive strains are limited to less than 10% and the

densification and hardening is insignificant at such small strain. The rate of term ̅ is

described as ̇ , the plastic multiplier:

̇̅ ̇ (4.14)

The plastic flow rule describes the evolution of . It is described as

̇ ̇ (4.15)

83
where sign is the signum function defined as

sgn(a)={ (4.16)

The plastic multiplier ̇ is a function describes how the rate of plastic strain varies

with the level of stress. There are many forms that are possible for ̇ . Based on the power

law form of the viscoplastic potential proposed by Perzyna [54] and Peric [55], it was

assumed that

| |
[( ) ]
̇ ̇̅ { (4.17)

where μ is the viscosity-related parameter, with dimension of time and β is a non-

dimensional rate-sensitivity parameter. Both μ and β are material constant and strictly

positive.

The mechanical analogs model is shown in Figure 4.2, where

(4.18)

Substituting Equation (4.18) into Equation (4.2)

= (4.19)

The stress-strain behavior then becomes

(4.20)

( ) ̇ (4.21)

84
Hence, the total stress became

( ) (4.22)

Figures 4.9 (a) and (b) are the plastic responses of PVC H100 in the out-of-plane

and in-plane directions. Both graphs indicate that the material exhibits viscoplasticity.

(a)

(b)
Figure 4.9 Experimental stress-strain curves showing viscoplasticity:
(a) Out-of-plane and (b) In-plane.

85
Equation (4.20) is rewritten as

́ (4.23)

where .

Substituting Equation (4.22) in Equation (4.17) and using that gives

( )
́ {[ ] } (4.24)

A MATLAB program, using ode45 solver, was written to curve fitting the plastic

response, and the results are shown as Table 4.4.

Table 4.4 Material properties for viscoplasticity.

Out-of-plane In-plane
β 0.0430 0.0606
μ (s) 27.5 E6 19.0 E6

Figure 4.10 (a) and (b) are the fitted curves for viscoplasticity. Table 4.5 provides

the comparison results of test data to the fitted curves in out-of-plane and in-plane

directions. More comparison details are shown in Appendix E. The difference between

test and predicted results for the out-of-plane direction is 5%. However, the difference for

the in-plane direction is up to 20% as the strain rates increase.

86
(a)

(b)

Figure 4.10 Fitted curves for viscoplastic reponse:


(a) Out-of-plane and (b) In-plane.

87
Table 4.5 Comparison results for viscoplasticity in out-of-plane and in-plane directions.

Out-of-plane In-plane

Strain Rate Test Flow Predicted Flow Diff. Test Flow Predicted Flow Diff.
(s-1) stress (psi) stress (psi) (%) stress (psi) stress (psi) (%)
0.0005 243.47 241.053 -0.99 149.97 143.30 -4.44
0.0050 265.85 266.147 0.11 171.76 164.24 -4.38
0.0500 283.87 293.145 3.27 178.56 188.76 5.71
0.5000 310.94 324.256 4.28 196.07 217.09 10.72
5.0000 339.33 358.182 5.56 209.70 249.63 19.04

4.1.3 Mullins Effect

As shown in Figure 4.11, the stress-strain curve reaches Point A at maximum

strain εmax. The coordinate position of Point A is (εmax, σys) and it becomes the beginning

point of the unloading. Before unloading, two values are defined: one is εv1, which is the

inelastic stain associated with εmax; the other is εp1, which is the plastic strain at εmax. The

values of εv1 and εp1, are given in Appendix F and were obtained from a MATLAB

program.

Figure 4.11 Stress-strain relations after plastic flow.

88
As reviewed in Chapter II, traditional methods to account for damage or so-called

Mullins softening are given by Simo [56] and Ogden-Roxburgh [45]. Both methods

derive the stress-strain relationship from strain energy density function. Because the

methods were initially intended to study the behavior of rubber-like materials under

cyclic loading, the Mullins effect was usually applied to materials under large

deformation, also called finite deformations. However, this research focuses on small

strain. Stress in finite deformations is usually correlated to the first and second Piola–

Kirchhoff stress tensors, the Biot stress tensor, or the Kirchhoff stress tensor. For small

strain, stress is related to the Cauchy stress tensor, which is the same as the second piola-

kirchhoff stress tensor.

There are differences of the studies of rubber-like materials and PVC foam.

Theories for rubber-like materials are derived for finite strain, while the theory for PVC

foam is derived for small strain. Both rubbers and foams will exhibit residual strains. The

residual strain effect in rubbers during Mullins damage is introduced by including another

potential damage function [2004]. Residual strains are already incorporated in the foam

during the primary loading phase.

Test data is used to determine a modified damage model to be appropriate for the

PVC foam. A damage parameter was defined as

(4.25)

89
-⁄
where , and and are material parameters. This function is

similar to Ogden-Roxburgh function with the exception that when because

residual strains and stresses are already incorporated in the primary loading curve.

From experimental observations and test data analysis, the parameters , and

η which were defined from the initial compressive stress-strain curves now become

functions of d. The permanent strain is also a function of d for later unloading and

reloading. Therefore, the mechanical analog shown in Figure 4.2 is modified as shown in

Figure 4.12. The bars over , , η and denote that these are damaged functions of

d.

Figure 4.12 Modified mechanical model.

90
From the mechanical model showed in Figure 4.2, the stress-strain relation with

damage is

̅ ( ̅ ) ̅ ( ̅ ) (4.26)

̅ ̅ ( ̅ ) (4.27)

where, ̅ , ̅ and ̅ are the damaged parameters.

For unloading,

̅ ̅ ̅
( ̅ ̅
) [ ̅
] ̅
̅ (4.28)

At the end of unloading, . Therefore,

̅ ̅ ̅
( ̅ ̅
) ̅ ̅̅̅̅̅
̅ (4.29)

The value of becomes the initial conditions for reloading.

For reloading,

̅ ̅ ̅
( ̅ ̅
) ( ̅
) ̅
̅ (4.30)

It is known that is damaged by the factor D:

̅ (4.31)

91
Figure 4.13 demonstrates how to get the plastic strain damage function ̅ .

Figure 4.13 Obtaining plastic strain damage function.

Based on Figure 4.13, the damaged plastic strain is

-
̅ - (4.32)

Both ̅ and ̅ are assumed as follows:

̅ { [ ] } (4.33)

̅ { [ ] } (4.34)

where and are two other material parameters, which could be evaluated by curve

fitting.

92
The OriginLab software program [34] was used for curve fitting. Figure 4.14

shows the experimental curve verses the simulation of Mullins damage at strain rate of

0.0005 s-1 in the out-of-plane direction. The priority of curve fitting is to match the shape

of hysteresis loop, catch the beginning and ending points of unloading and reloading, and

simulate the energy dissipation as accurate as possible.

Figure 4.14 The experimental curve verses the simulation of Mullins damage at
strain rate of 0.0005 s-1 in the out-of-plane direction.

After curve fitting, the proper value of r, m, , and are found as, r=1.29,

m=0.035, C1=2 and C2=220. The values of r, m, and are material properties and

are therefore constants for one particular material. By keeping these 4 parameters the

same, but changing the strain rates, the stress-strain curves are shown as in Figure 4.15. A

comparison of the hysteresis at different strain rates in the out-of-plane direction indicates

that the method of applying Mullins effect to simulate the damage in polymer foam failed

93
to capture all the strain rate effect, as shown in Figure 4.15 (a)-(d). Furthermore, given a

set of values for r, m, , and cannot match the test data for different strain amplitudes

at a fixed strain rate.

94
(a) (b)

(c) (d)

(e)
Figure 4.15 Comparison of the hysteresis at different strain rates in out-of-plane direction
with r=1.29, m=0.035, C1=2 and C2=220: (a) 0.0005 s-1, (b) 0.005 s-1, (c) 0.05 s-1, (d)
0.5 s-1and (e) 5 s-1.

95
(a) (b)

(c) (d)

Figure 4.16 Comparison of the hysteresis at different strain amplitude in out-of-plane


direction with r=1.29, m=0.035, C1=2 and C2=220: (a) 10%, (b) 8%, (c) 6% and (d) 4%.

96
4.2 Viscoelastic Damage Mechanics Model

In the viscoelastic damage mechanics model, the undamaged and damaged

behavior are distinguished by a damage initiation criterion. The damage initiation

criterion for the foam is associated with the onset of buckling or what was described as

plastic behavior in the previous Mullins model.

A simple damage initiation criterion for the foam is proposed as

(3.35)

The experiments indicate that for all the test strain rates, =0.023 for the out-of-plane

compression mode. See Figure 4.17, for example.

Figure 4.17 Damage initiations at , using actual test for out-of-plane compression.

The stress-strain response before damage at εcr is the same viscoelastic response

discussed in Section 4.1.1. In this new damage model, it is assumed that the damage

occurs right after strain exceeds εcr and the damage is controlled by the maximum strain

97
(εmax) and the history of loading or the flow stress ( ). Few things should be clarified,

such as how to determine the flow stress , and how to apply damage in the model. The

flow stress is determined at the end of viscoelastic response and just before damage

begins. As indicated in Figure 4.17, increases with increasing strain rate. This is also

shown in Figure 4.18 (a), where the stain rate effect is synonymous with flow stress or

history effect. The unloading curves for both strain rates are different because damage is

different. As the maximum strain increases at a fixed stain rate, the damage changes,

however. In general, there is more hysteresis with a larger maximum strain as explained

in Figure 4.18 (b), and observed in test results.

98
(a)

(b)

Figure 4.18 Stress-strain response using damage mechanics:


(a) History effect and (b) Effect of maximum strain amplitude. .

The critical strain at damage initiation from tests should not be used as the

predictive damage initiation criterion in the model because the previous predicted

response would not perfectly match test results. As shown in Figure 4.19, the dashed line

indicates that for different strain rates, the damage should initiate at a strain of 0.019 in

order to get similar flow stress as in the experiments. This is somewhat lower than the

critical strain from the test (see Fig. 4.17), but this is because the predicted viscoelastic

response yields higher stresses than the test. Once damage initiates, the flow stress is

constant with strain amplitude and only depends on the strain rate or previous history of
99
loading. For various strain rates, flow stresses are determined and these are given in

Table 4.1.

Figure 4.19 Determining flow stress at , using simulated viscoelastic response for
out-of-plane compression.

Table 4.6 Flow stresses at various strain rates in out-of-plane compression.

Strain rate
(s-1) 0.0005 0.005 0.05 0.5 5

Plastic stress
(psi) 245 264 287 305 325

The mechanical model for using viscoelastic damage mechanics is summarized in

Figures 4.20 (a) and (b). All the material parameters: , and η are modified with

the damage as indicated by the “bar”. From former discussion, it is known that all the

damaged parameters will be functions of the combination of and the history of the

material behavior before damage initiates at different strain rates.

100
(a)

(b)

Figure 4.20 Mechanical analog for a damage mechanics model:


(a) Before damage and (b) After damage.

From the mechanical model showed in Figure 4.20 (b), the stress-strain relation

with damage is

̅ ̅ ( -̅ ) (4.41)
The overstress is

̅ ̅ ̅ ̅ ̇ (4.42)

Solving for from the above equation gives

̅ ̅ ( - ) ̅
(̅ - -̅ ) [ ̅
] ̅
(4.43)

101
where, ̅ is the plastic strain at . This plastic strain depends on the damaged

material properties, which in turn depends on and . At the end of unloading,

, and becomes

̅ ̅ ̅
̅ (̅ - -̅ ) - ̅ ̅
(4.44)

The above equation is the initial condition for on reloading. For reloading,

̅ ̅ ̅
(̅ ̅
) (- ̅
) -̅ (4.45)

The flow stress is

̅ ̅ ( -̅ ) (4.46)

Therefore, ̅ could be determined by


̅ (4.47)

By curve fitting, the functions for the damaged ̅ , ̅ and ̅ in out-of-plane

compression are found as follows:

̅ -
(4.48)

̅ -0.5469) - +0.1278 (4.49)

̅ [ ( ) ]

(4.50)

102
Substituting Equations (4.48) and (4.49) into Equation (4.47) gives

-
-
̅ (4.50)
- - -

This new damage model can successfully describe both the strain rate and strain

amplitude effects found in the out-of-plane compression tests. The predicted curves are

comparable with test data at strain amplitude of 10% in the out-of-plane direction and

various strain rates. Results from Figures 4.21 (a)-(e) for strain rates 0.0005 s-1, 0.05 s-1

and 5 s-1 are superimposed in Figure 4.22 to show that the effect of increasing strain rate

at a fixed strain amplitude is captured with this new constitutive model.

The new damage model does not only capture strain rate effects but also the strain

amplitude effects. In Figure 4.23, predicted hysteresis at a fixed strain rate of 0.0005 s-1

and increasing strain amplitude or are compared with the out-of-plane test results.

Good agreement was found between the two, so that this new viscoelastic damage model

can simulate the correct response with increasing strain amplitude and fixed strain rate.

More comparisons of the simulation curves and experimental results are shown in

Appendix G.

103
(a) (b)

(c) (d)

(e)
Figure 4.21 Comparison of simulation curves with experimental data at strain amplitude
of 10% in the out-of-plane direction: (a) 0.0005 s-1, (b) 0.005 s-1, (c) 0.05 s-1, (d) 0.5 s-
1
and (d) 5 s-1.

104
Figure 4.22 Comparison of simulation curve and test data at various strain rates and fixed
strain amplitude of 10% in the out-of-plane direction.

Figure 4.23 Comparison of simulation curve and test data at various strain amplitudes and
fixed strain rate of 0.0005 s-1 in the out-of-plane direction.

105
CHAPTER V

CONCLUSIONS AND FUTURE WORK

The purpose of this research was to develop a phenomenological-based,

constitutive model for polymer foams under cyclic loading. Monotonic and cyclic

compression and pure shear tests were done on Divnycell H100 foam to produce

hysteresis data in both in-plane and out-of-plane directions. The strain rates ranged from

0.0005 s-1 to 5 s-1. The compressive strain amplitudes were limited to less than 10%,

while the engineering shear strain amplitudes were less than 20%. Experimental data

indicates that PVC H100 foam displays transversely isotropic behavior under both

compression and pure shear modes.

Two one-dimensional constitutive models were developed from the test data. The

first constitutive model is related to current crushable foam model which is based on

metallic foams. In this research, the model consisted of an equilibrium spring in parallel

with a Maxwell element and in series a Prandtl element. Damage functions were

introduced for permanent crushing of cells (plastic deformation) and progressive damage

of cells (Mullins effect). This constitutive model was able to capture yielding,

viscoelastic and viscoplastic response, as well as hysteresis of the foam, but it failed to

106
describe the strain-rate dependency that was found in the experiments. The second

constitutive model is based on damage initiation and evolution of the foam. It consisted

of an equilibrium spring in parallel with a Maxwell element. After damage initiation,

damaged spring and dashpot functions were found. The damaged functions were found to

depend on the magnitude and history of the deformation. Good agreement with out-of-

plane compression test results was reported. Similar modeling could be used for in-plane

compression as well as shear modes.

The ideas developed in this research provide a foundation for three-dimensional

constitutive modeling of polymer foams. The present study only addresses a one-

dimensional constitutive model for the foam. Future work must be done to examine the

multi-axial stress-strain response of foam. This would involve running experiments on

the foam in multiaxial stress states and under cyclic loading. In addition to this, the

proposed model is limited to material behavior under a constant strain rate. In an actual

impact or dynamic problem the strain rates and strain at point may change in time. This

true history effect in the foam material has not been addressed, and would require further

experimental tests to be done.

107
REFERENCES

[1] D. Zenkert. (1997) The handbook of sandwich construction, The Chameleon Press
Ltd, London, United Kingdom.

[2] DIAB. (2011) http://www.diabgroup.com/europe/concept/e_concept_1.html.

[3] J. Reilly. (2007) PVC products- technical details. Foamalite Ltd.


http://www.gilmanbrothers.com/foamalitetechmanual.pdf

[4] D.H. Pascoe. High tech materials in boat building, D. H. PASCOE & CO., Inc.
http://www.yachtsurvey.com/HiTech.htm

[5] L. Aktay, A.F. Johnson, and M. Holzapfel. (2005) Prediction of impact damage on
sandwich composite panels, Comput Mater Sci, 32: 252-60.

[6] S. Abrate. (1998) Impact on composite structures. Cambridge University Press, UK.

[7] U. Icardi, and L. Ferrero. (2009) Impact analysis of sandwich composites based on a
refined plate element with strain energy updating, Compos Struct, 89:35-51.

[8] S. Ryan, F. Schaefer, and W. Riedel. (2006) Numerical simulation of hypervelocity


impact on CFRP/Al HC SP spacecraft structures causing penetration and fragment
ejection, Int J Impact Eng, 33: 703-12.

[9] V.S. Deshpande, and N.A. Fleck. (2000) Isotropic constitutive model for metallic
foams, Journal of the Mechanics and Physics of Solids, 48: 1253-1276.

[10] V.S. Deshpande, and N.A. Fleck. (2001) Multi-Axial Yield Behavior of Polymer
Foams, Acta Mater, 49: 1859-1866.

108
[11] ABAQUS, Inc. (1997) ABAQUS User Manual, Version 6.9.

[12] L.J. Gibson, and M.F. Ashby. (1982) The mechanics of three-dimensional cellular
materials, Proceedings of the Royal Society, London, 382: 43-59.

[13] L.J. Gibson, M.F. Ashby, G.S. Schajer, C.I. Robertson. (1982) The mechanics of
two-dimensional cellular materials, Proceedings of the Royal Society, London, 382: 25-
42.

[14] R. Hill. (1978) Aspects of invariance in solid mechanics, Advances in Applied


Mechanics, 18: 1-78.

[15] B. Storåkers. (1986) On material representation and constitutive branching in finite


compressible elasticity, Journal of the Mechanics and Physics of Solids, 34(2): 125-145.

[16] P.J. Blatz, and W.L. Ko. (1962) Application of finite elasticity to the deformation of
rubbery materials, Transactions of the Society of Rheology, 6: 223-251.

[17] R.W. Ogden. (1972) Large deformation isotropic elasticity – on the correlation of
theory and experiment for compressible rubberlike solids, Proceedings of the Royal
Society, London, 328: 567-583.

[18] R.W. Ogden, and D.G. Roxburgh. (1999) A pseudo-elastic model for the mullins
effect in filled rubber, Proceedings of the Royal Society, London, 455:2861-2877.

[19] M. Sadighi, and H. Pouriayevali. (2008) Quasi-static and low-velocity impact


response of fully backed or simply supported sandwich beams, J Sand Struct Mater, 10:
499-524.

[20] I. Ivañez, C. Santiuste, and S. Sanchez-Saez. (2010) FEM analysis of dynamic


flexural behavior of composite sandwich beams with foam core, Compos Struct, 92:
2285-91.

109
[21] M.S. Hoo Fatt, H. Surabhi, and Y. Gao. (2010) Blast response of sandwich shells
with crushable foam cores, in the Proceedings of the IMPLAST, Providence, RI, October
12-14, 2010.

[22] M.S. Hoo Fatt, and L. Palla. (2009) Analytical modeling of composite sandwich
panels under blast loads, Journal of Sandwich Structures and Materials, 11: 357-38.

[23] V.L.Tagarielli, V.S. Deshpande, and NA. Fleck. (2010) Prediction of the dynamic
response of composite sandwich beams under shock loading, International Journal of
Impact Engineering, 37: 854-864.

[24] D. Balkan, and Z. Mecitoglu. (2009) Dynamic response of sandwich plate with
viscoelastic core under blast load, in the proceedings of 7th EUROMECH Solid
Mechanics Conference, Lisbon, Portugal, September 7-11, 2009.

[25] Z. Xue, and J.W. Hutchinson. (2003) Preliminary assessment of sandwich plates
subjected to blast loads, International Journal of Mechanical Sciences, 45: 687-705.

[26] E. E. Gdoutos, I. M. Daniel, and K. A. Wang. (2002) Failure of cellular foams under
multiaxial loading, Applied Science and Manufacturing, 33(2):63-176.

[27] M. D. Isaac, J.J. Luo, and M.S. Patrick. (2007) Mechanical and failure
characterization of textile composites, in the Proceedings of 16th International
Conference on Composite Materials, Kyoto, Japan, July 8-13, 2007.

[28] F. Pan. (2008) Modeling of the Rate Responsive Behavior of Elastomer Foam
Materials, Journal of Engineering Materials and Technology, 130(1):1-6.

[29] F. S. Chang. (1995) Constitutive equation development of foam materials, Ph.D.


thesis, Wayne State University, Detroit, MI.

[30] F. Pan, S. Quander, and A. Ma. (2005) Use of material models for elastomer foams
in ABAQUS and LS-DYNA, in the Proceedings of the Fourth European Conference for
Constitutive Models for Rubber (ECCMR), 593-597, Stockholm, June 27-29, 2005.

110
[31] U. E. Ozturka, and G.Anlasb. (2011) Finite element analysis of expanded
polystyrene foam under multiple compressive loading and unloading, Materials &
Design, 32(2):773-780.

[32] P. M. Schubel, J. Luo, and I. M. Daniel. (2005) Low velocity impact behavior of
composite sandwich panels, Composites, 36:1389-1396.

[33] V.I. Rizov. (2006) Non-linear indentation behavior of foam core sandwich
composite materials-A 2D approach, Computational Materials Science, 35(2): 107-115.

[34] I. Ivañez, C. Santiuste, S. Sanchez-Saez. (2010) FEM analysis of dynamic flexural


behavior of composite sandwich beams with foam core, Compos Struct, 92:2285-91.

[35] J.P. Hou, N. Petrinic, C. Ruiz, SR. Hallett. (2000) Prediction of impact damage in
composite Plates, Compos Sci Technol, 60(2):273–328.

[36] V.S. Deshpande, and N.A. Fleck. (2001) Multi-axial yield behaviour of polymer
foams, Acta Mater, 49:1859-66.

[37] V.S. Deshpande, and N.A. Fleck. (2000), Isotropic constitutive models for metallic
foams, Journal of the Mechanics and Physics of Solids, 48(6-7):1253-1283

[38] G. Vladimír. (2010) Testing and Application of New Phenomenological Material


Model for Foam Materials, http://www.posterus.sk/?p=392.

[39] G. Holzapfel. and J. Simo. (1996) A new viscoelastic constitutive model for
continuous media at finite thermomechanical changes, Int. Solids Struct, 33: 3019-3034.

[40] K. Ti, B.B.K. Huat, J. Noorzaei, M. Jaafar and G.S. Sew (2009) A review of basic
soil constitutive models for geotechnical application, Electronic Journal of Geotechnical
Engineering, 14:1-17.

[41] H.S. Yu (2006) Plasticity and Geotechnics, Advances in Mechanics and


Mathematics, Volume 13, First Edition, Springer, Berlin, Germany.

111
[42] K.K. Chawla (2008) Mechanical Behavior of Materials, second edition, Cambridge
University Press, United Kingdom.

[43] D.P. Zekkos, J.D. Bray, and M.F. Riemer. (2006) Shear modulous reduction and
material damping relations for municipal solid-waste, in the Proceedings of the 8th U.S.
National Conference on Earthquake Engineering, San Francisco, California, USA, Apirl
18-22, 2006.

[44] Y. Wang and T.D. O’Rourke. (2007) Interpretation of secant shear modulus
degradation characteristics from pressuremeter tests, Journal of Geotechnical &
Geoenvironmental Engineering, 133(12): 1556-1566.

[45] R.W. Ogden and D.G. Roxburgh. (1999) A pseudo-elastic model for the Mullins
effect in filled rubber, Mechanical Physics and Engineering Sciences, 455: 2861-2877.

[46] J.C. Simo, and J.W. Ju. (1987) Strain and stress based continuum damage models,
Part II: Computational Aspects”, Int. J. Solids & Struct., 23(7):841-869.

[47] S. Yazdani, and H.L. Schreyer. (2003) Nonlinear response of plain concrete shear
walls with damage, International Journal of IT in Architecture, 1(3): 251-258.

[48] N. R. Hansen. and H.L. Schreyer. (1994) A thermodyamically consistent framework


for theories of elastoplasticity coupled with damage, Int. J. Solid Structures, (3):359-389.

[49] R.D. Borst, P. Jerzy, and G.D. Marc. (1999) Geers On coupled gradient-dependent
plasticity and damage theories with a view to localization analysis, Eur. J. Mech.
A/Solids, 18: 939-962.

[50] DIAB (2005)


http://www.diabgroup.com/aao/a_literature/a_pdf_files/R_Grade_E.pdf.

[51] OriginLab, Version7.5. (2003)


http://www.originlab.com/index.aspx?go=Support&pid=

112
[52] A.F. Bower. (2008) Applied Mechanics of Solids, http://solidmechanics.org/.

[53] De Souza Neto, E.A., Peri´c, D. and Owen, D.R.J. (2008) Computational Methods
for Plasticity: Theory and Applications, John Wiley & Sons, Inc., Chichester, U.K.

[54] P. Perzyna. (1963) The constitutive equations for rate sensitive plastic materials,
Quart. Appl. Math., 20:321-332.

[55] De Souza Neto, E.A., Peri´c, D. and Owen, D.R.J. 1993. Some Aspects of
Formulation and Implementation of Ductile Damage at Finite Strains, ASME 93
Conference: Computational Mechanics in the U.K.

[56] J.C.Simo (1987) On a fully three-dimensional finite-strain viscoelastic damage


model: formulation and computational aspects, Computational Methods of Applied
Mechanics and Engineering, 60: 153-173.

[57] Mullins, L., 1947. Effect of Stretching on the Properties of Rubber. Journal of
Rubber Research 16, 275-289.

[58] Dorfmann, A., Ogden, R.W., 2004. A constitutive model for the Mullins effect with
permanent set in particle-reinforced rubber. International Journal of Solids and
Structures 41, 1855–1878.

[59] Henkel, Technical Data Sheet, www.henkel.com/tds-search-29668.htm.

[60] ASTM C273 (2007): Standard test method for shear properties of sandwich core
materials.

[61] ISO 1922 (2001): Rigid cellular plastics –Determination of shear strength.

113
APPENDIX A

UNIAXIAL COMPRESSION TEST RESULTS

114
(a) (b)

(c) (d)

Figure A.1 Out-of-plane monotonic test under various strain rates.

115
APPENDIX B

CYCLIC COMPRESSION TEST RESULTS

116
(a) (b)

(c) (d)

(e)

Figure B.1 Strain Amplitude Effect (out-of-plane).

117
(a) (b)

(c) (d)

(e)

Figure B.2 Strain Rate Effect (out-of-plane).

118
(a) (b)

(c) (d)

(e)

Figure B.3 Strain Amplitude Effect (in-plane).


119
(a) (b)

(c) (d)

(e)

Figure B.4 Strain Rate Effect (in-plane).


120
APPENDIX C

CYCLIC SHEAR TEST RESULTS

121
(a) (b)

(c) (d)

(e)

Figure C.1 Strain Amplitude Effect (out-of-plane).


122
(a) (b)

(c) (d)

(e)

Figure C.2 Strain Rate Effect (out-of-plane).

123
(a) (b)

(c) (d)

(e)

Figure C.3 Strain Amplitude Effect (in-plane).


124
(a) (b)

(c) (d)

(e)

Figure C.4 Strain Amplitude Effect (in-plane).


125
APPENDIX D

COMPARISON OF VISCOELASTIC CURVES

126
(a) (b)

(c) (d)

(e)

Figure D.1 Viscoelastic curves (out-of-plane).


127
(a) (b)

(c) (d)

(e)

Figure D.2 Viscoelastic curves (in-plane).

128
APPENDIX E

COMPARISON OF VISCOPLASTIC CURVES

129
(a) (b)

(c) (d)

(e)

Figure E.1 Viscoplastic curves (out-of-plane).


130
(a) (b)

(c) (d)

(e)

Figure E.2 Viscoplastic curves (in-plane).


131
APPENDIX F

INELASTIC STRAIN VALUES

Table F.1 Out-of-plane.

Strain rate

0.0005 s-1 0.0050 s-1 0.0500 s-1 0.5000 s-1 5.0000 s-1

0.04 0.01729 0.02271 0.01909 0.02124 0.01849 0.01932 0.00429 0.01860 0.00052 0.01703

0.06 0.01729 0.04273 0.01909 0.04085 0.02054 0.03903 0.00711 0.03855 0.00094 0.03859

0.08 0.01729 0.06270 0.01909 0.06062 0.02097 0.05897 0.00961 0.05956 0.00132 0.05766

0.10 0.01729 0.08271 0.01909 0.08091 0.02106 0.07892 0.01153 0.07840 0.00170 0.07770

Table F.2 In-plane.

Strain rate

0.0005 s-1 0.0050 s-1 0.0500 s-1 0.5000 s-1 5.0000 s-1

0.04 0.01870 0.02093 0.02041 0.01928 0.00818 0.01728 0.00108 0.01429 0.00013 0.01090

0.06 0.01870 0.04134 0.02073 0.03928 0.01229 0.03608 0.00196 0.03548 0.00024 0.03157

0.08 0.01870 0.06122 0.02075 0.05938 0.01557 0.05683 0.00270 0.05388 0.00033 0.05047

0.10 0.01870 0.08130 0.02075 0.07925 0.01782 0.07658 0.00349 0.07426 0.00043 0.07066

132
APPENDIX G

COMPARISON OF PREDICTED RESPONSE TO TEST RESULTS FOR OUT-OF-

PLANE COMPRESSION USING VISCOELASTIC DAMAGE MEDEL

133
(a) (b)

(c) (d)

(e)

Figure G.1 Strain amplitude=8% (out-of-plane).


134
(a) (b)

(c) (d)

(e)

Figure G.2 Strain amplitude=6% (out-of-plane).


135
(a) (b)

(c) (d)

Figure G.3 Strain amplitude=4% (out-of-plane).

136
(a) (b)

(c)

Figure G.4 Strain amplitude=2% (out-of-plane).

137

You might also like