experimental characterization and modeling of high

194
EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH STRENGTH MARTENSITIC STEELS BASED ON A NEW DISTORTIONAL HARDENING MODEL BY ELIZABETH K BARTLETT A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY UNIVERSITY OF FLORIDA 2018

Upload: others

Post on 18-Mar-2022

7 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH STRENGTH MARTENSITIC STEELS BASED ON A NEW

DISTORTIONAL HARDENING MODEL

BY

ELIZABETH K BARTLETT

A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT

OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY

UNIVERSITY OF FLORIDA

2018

Page 2: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

2

© 2018 Elizabeth K Bartlett

Page 3: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

3

To my daughter, Gwyneth Webb, who helps me to have fun, laugh, and love through the most difficult times in my life.

Page 4: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

4

ACKNOWLEDGEMENTS

First, I would like to thank all the members of my supervisory committee

for their support. I would especially like to thank my committee chair, Prof. Oana

Cazacu, for her support and constructive criticisms. It was an honor and a

privilege to study her research. I would also like to thank Dr. Benoit Revil-

Baudard for his expertise in numerical methods.

I am so grateful for the support of the 96th Test Wing, my sponsoring

facility in the Science, Mathematics, and Research for Transformation (SMART)

scholarship program. Specifically, I would like to thank my commanders and

supervisor, Mr. Ron Lutz, as well as, Mrs. Linda Busch and Dr. Betta Jerome.

I also greatly appreciate the support of the Air Force Research Laboratory

(AFRL) in the experimental characterization of Eglin steel: Dr. Geremy Kleiser

and Dr. Philip Flater for their training in quasi-static testing, Dr. Brad Martin and

Dr. Xu Nie for their expertise in dynamic split Hopkinson pressure bar (SHPB)

testing, Dr. Rachel Abrahams and Dr. Sean Gibbons for their instruction in

microscopy and material science, and Mr. Richard Harris for his expertise in

material characterization at the Advanced Weapons Effects Facility (AWEF).

I extend my gratitude to my friends and family for listening to me ramble

on and on about subsequent yield surfaces and various hardening models.

Finally, I would like to thank Gwyneth Webb, my 12-year old daughter, for her

encouragement and patience over long nights and weekends in room 171 of the

University of Florida (UF) Research and Engineering Education Facility (REEF).

Page 5: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

5

TABLE OF CONTENTS

PAGE

ACKNOWLEDGEMENTS ..................................................................................... 4

LIST OF TABLES .................................................................................................... 7

LIST OF FIGURES .................................................................................................. 8

LIST OF ABBREVIATIONS .................................................................................... 13

ABSTRACT ......................................................................................................... 16

CHAPTER

1 INTRODUCTION ............................................................................................. 18

1.1 Background of Ultra High Strength Martensitic Steel..................... 18 1.2 Survey of the Experimental Studies on Martensitic Steels ............ 23 1.3 Eglin Steel, ES-1 ........................................................................... 32 1.4 Elastic-Plastic Modeling ................................................................. 38 1.5 Goals of Current Research ............................................................ 48

2 MICROSCOPY ................................................................................................ 50

2.1 Optical Microscopy of ES-1 ........................................................... 51 2.2 Material Characterization using SEM ............................................ 55

3 MECHANICAL CHARACTERIZATION OF ES-1 ......................................... 63

3.1 Hardness of Eglin Steel ................................................................. 63 3.2 Quasi-Static Mechanical Characterization ..................................... 65 3.3 Dynamic Experimental Characterization of Eglin Steel .................. 82 3.4 Cyclic Experimental Characterization of ES-1 ............................... 99 3.5 Summary of the Experimental Characterization of ES-1 ............. 104

4 ELASTIC-PLASTIC MODEL FOR EGLIN STEEL ..................................... 105

4.1 Development of the Yield Criterion .............................................. 105 4.2 Asymmetric Hardening ................................................................ 112 4.3 Implementing the Proposed Yield Function ................................. 137 4.4 Implications of the Proposed Yield Function ............................... 143

Page 6: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

6

5 FINITE ELEMENT ANALYSIS ..................................................................... 145

5.1 Review of Finite Element Analysis .............................................. 145 5.2 Implementation of the Elastic-Plastic Model in FEA .................... 149 5.3 Finite Element Analysis of Ultra High Strength Martensitic Steel 151 5.4 Elastic-Plastic Model Predictions ................................................ 170

6 CONCLUSIONS AND RECOMMENDATIONS .......................................... 178

REFERENCES ................................................................................................. 184

BIOGRAPHICAL SKETCH ............................................................................... 194

Page 7: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

7

LIST OF TABLES

Table Page

1-1 Monotonic and cyclic material properties. ................................................ 30

1-2 Chemical composition by percent weight of AF-1410, SAE 4340, and ES-1. .............................................................................................................. 33

1-3 Mechanical properties under uniaxial, quasi-static tensile loading of AF-1410, SAE 4340, and ES-1. ..................................................................... 34

2-1 Heat treatment schedule for ES-1. ........................................................... 50

2-2 ES-1 grinding and polishing schedule. ..................................................... 51

2-3 Chemical composition of ES-1 for EDS. ................................................... 56

3-1 Eglin steel material characterization test matrix. ...................................... 63

3-2 Yield stress and quasi-static compression test data. ................................ 70

3-3 Coefficients involved in Swift and Voce Hardening Laws for ES-1. .......... 70

3-4 Summary of quasi-static tension test data................................................ 72

3-5 Hardening law parameters tensile round specimens. ............................... 73

3-6 Tension-compression asymmetry ratio of ES-1 by plastic strain. ............. 76

3-7 Hardening law parameters flat tension specimens. .................................. 77

3-8 Johnson and Cook material constants for forged, cast, and cast and HIP’d ES-1. ........................................................................................................ 94

4-1 Cazacu and Barlat yield criterion parameters. ........................................ 140

4-2 Points for linear interpolation of the Cazacu and Barlat asymmetry parameter, c. .......................................................................................... 141

5-1 Finite element analysis simulation matrix. .............................................. 152

Page 8: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

8

LIST OF FIGURES Figure Page

1-1 Tensile yield stress vs. strain-to-failure of current steels and expected strengths of third-generation UHSS. ........................................................ 19

1-2 The unit cell of single-crystal microstructures of solid steel. ..................... 20

1-3 Illustration of plastic deformation mechanisms. ........................................ 22

1-4 The fractional strength differential parameter vs. the carbon content by percent weight for different SAE 4300 series. .......................................... 24

1-5 The effect of tempering temperature on the strength differential of quenched and tempered 4340 UHSS. ..................................................... 25

1-6 The effect of tempering on the flow stress of various steels. .................... 28

1-7 Absolute flow stress vs. plastic strain of SAE 4340 from hysteresis loops of cyclic testing. ........................................................................................ 31

1-8 STF vs. tensile yield stress for different steels including conventional steels and UHSS. ..................................................................................... 35

1-9 ES-1 symmetric plate impact experimental setup of Martin et al. ............. 37

1-10 Tresca yield surface in Haigh-Westergaard space with the hydrostatic axis and deviatoric plane. ......................................................................... 40

1-11 Von Mises yield surface in Haigh-Westergaard space with the longitudinal axis as the hydrostatic axis. ..................................................................... 41

1-12 Projection in the biaxial plane of the yield surfaces of Drucker, Mises, and Tresca yield criteria. ................................................................................. 42

1-13 Cazacu and Barlat yield surface in Haigh-Westergaard space. ............... 44

1-14 Projection in the biaxial plane of the Cazacu and Barlat yield surface corresponding to flow stress ratios of 3/4, 1, and 5/4. .............................. 45

2-1 Z-stack of several pores on the polished surface of cast eglin steel specimen using CDIC. ............................................................................. 52

2-2 Photomontage stitched from nine individual images of the polished surface of cast ES-1 specimen containing several pores. ..................................... 53

2-3 Polished surface of ES-1 materials at 200 times magnification in an optical microscope with brightfield illumination. ................................................... 54

Page 9: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

9

2-4 Nitol etched cast material surface at 100 times magnification in Keyence optical microscope under brightfield illumination. ..................................... 55

2-5 Orientation map of forged Eglin steel demonstrating prior austenite grain boundaries and martensitic lath structure.Figure 2-6. Orientation map of forged Eglin steel demonstrating prior austenite grain boundaries and martensitic lath structure. ......................................................................... 57

2-6 Orientation map of cast and HIPd Eglin steel demonstrating prior austenite grain boundaries and martensitic lath structure. ....................................... 58

2-7 Orientation map of cast Eglin steel demonstrating prior austenite grain boundaries and martensitic lath structure................................................. 59

2-8 Pole figures of (001), (101), and (111) for forged, cast and HIP’d, and cast Eglin steel. ................................................................................................ 60

3-1 Buehler Digital Hardness Tester MMT-3.Figure 3-1. Buehler Digital Hardness Tester MMT-3. ......................................................................... 64

3-2 AWEF quasi-static experimental characterization setup .......................... 65

3-3 Quasi -static cylindrical compression test specimens. ............................. 67

3-4 Quasi-static round bar compression test stress-strain results .................. 68

3-5 Comparison of the quasi-static compressive stress-strain response for the ES-1 materials. ......................................................................................... 69

3-6 Hardening of ES-1 under compression according to Swift and Voce ....... 71

3-7 Schematic of the quasi-static round tensile test specimen. ...................... 72

3-8 Stress strain curves for ES-1 quasi-static tension test results of forged, cast and HIP’d, and cast no HIP specimens. ........................................... 74

3-9 Fracture surface of the round specimens following tensile tests .............. 75

3-10 Fracture surface of the round specimens following tensile tests .............. 76

3-11 Schematic of the quasi-static (flat) pin-loaded tensile test specimen. ...... 78

3-12 Quasi-static tension characterization stress-strain response. .................. 79

3-13 Camera layout to capture 2-D measurements of the strain evolution on the face and the side of the flat specimens during quasi-static tensile testing. ................................................................................................................. 81

3-14 Width and thickness strain of flat specimens under uniaxial tension ........ 82

Page 10: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

10

3-15 Schematic SHPB with an illustration of the propagating strain waves. ..... 83

3-16 A free body diagram of a portion of the SHPB of length dx and cross sectional area A. ...................................................................................... 87

3-17 Raw data from the dynamic SHPB characterization ................................. 88

3-18 The SHPB test specimen subjected to forces at the incident and transmitted bar interfaces. ........................................................................ 89

3-19 UF REEF test equipment for dynamic SHPB characterizations ............... 90

3-20 Strain waves at the incident and transmitted strain gages ....................... 91

3-21 Stress strain response of forged, cast and HIP’d, and cast no HIP ES-1 during SHPB tests. ................................................................................... 92

3-22 Forged Eglin steel cylindrical SHPB test specimens ................................ 93

3-23 Comparison of quasi-static and dynamic stress-strain response ............. 95

3-24 Initial yield stress-strain-rate dependence of AF-1410, AISI 4340, and ES-1 observed by Last in comparison to the cast material. ........................... 96

3-25 Strain rate in forged, cast and HIP’d, and cast no HIP ES-1 during SHPB tests. ........................................................................................................ 97

3-26 Dynamic pulse shapers made of copper, steel, and Teflon ...................... 98

3-27 Strain rate of forged test specimens without a pulse shaper, with the first pulse shaper, and with the second pulse shaper. .................................... 99

3-28 Stress strain curves of forged specimens without a pulse shaper, with the first pulse shaper design and the second pulse shaper design. ............. 100

3-29 UF REEF materials lab for cyclic characterization of ES-1 .................... 101

3-30 Hysteresis loops of ES-1 under completely reversed displacement ....... 102

3-31 Uniaxial absolute flow stress of ES-1 at the limits of the elastic region under cyclic loading by plastic strain. ..................................................... 103

3-32 Monotonic and kinematic stress-strain curves of Eglin steel .................. 105

4-1 Illustration of properties of yield surfaces with an associated flow rule .. 109

4-2 Cazacu and Barlat initial and two subsequent yield surfaces using classical isotropic hardening of forged ES-1 in the deviatoric plane. ...... 113

Page 11: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

11

4-3 The first stress-strain cycle of forged ES-1 during cyclic testing at completely reversed displacement of 0.57, 0.62 and 0.67 mm and the uniaxial tensile and compressive flow stress. ......................................... 118

4-4 Flow stress at the limits of the elastic region by plastic strain in forged ES-1 under cyclic loading. ............................................................................ 119

4-5 The uniaxial back stress as a function of plastic strain for forged ES-1 under cyclic loading. ............................................................................... 120

4-6 Subsequent yield surfaces using classical isotropic, redefined isotropic, and kinematic hardening under uniaxial tensile loading. ........................ 121

4-7 Evolution of the material asymmetry parameter, c, for forged ES-1 using the limits of the elastic region under cyclic loading. ................................ 122

4-8 Uniaxial flow stress at the limits of the elastic region and yield surfaces in the deviatoric plane for several values of plastic strain in forged ES-1 under cyclic loading. ............................................................................... 123

4-9 Cazacu and Barlat subsequent yield surface in the deviatoric plane following uniaxial tension, compression, or pure shear strain. ............... 125

4-10 Decomposition of the yield surface with distortional hardening .............. 126

4-11 Cazacu and Barlat yield surface under both uniaxial tension and compression and the common component............................................. 128

4-12 Initial and subsequent Cazacu and Barlat yield surface under pure shear strain. ..................................................................................................... 129

4-13 Cazacu and Barlat initial and third invariant component of the subsequent yield surfaces using asymmetric isotropic hardening under uniaxial tensile loading in the deviatoric plane. ............................................................... 136

4-14 Cazacu and Barlat yield surface in the deviatoric plane under uniaxial tension, uniaxial compression, and pure shear using distortional hardening. .............................................................................................. 138

4-15 Cazacu and Barlat material parameter, c, by plastic strain for forged specimens, cast and HIP’d specimens and cast specimens. ................. 139

4-16 Cazacu and Barlat yield surfaces in the deviatoric plane ....................... 142

5-1 Stress distribution according to the model within forged ES-1 under quasi-static uniaxial tension. ............................................................................ 153

5-2 Stress and plastic strain for forged ES-1. ............................................... 154

Page 12: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

12

5-3 The longitudinal strain distribution from DIC and FEA forged specimens. ............................................................................................................... 155

5-4 The longitudinal strain distribution from DIC and FEA cast and HIP’d specimens. ............................................................................................. 155

5-5 The longitudinal strain distribution from DIC and FEA of cast specimens. ............................................................................................................... 156

5-6 Stress-strain curves of flat DIC and FEA specimens in tensile loading .. 157

5-7 The uniform, longitudinal stress distribution within the forged round finite element specimens under quasi-static compressive loading ................. 158

5-8 Stress-strain curves of round DIC and FEA compression specimens .... 159

5-9 The experimental and model load-displacement curves ........................ 162

5-10 Stress-strain curves for round DIC and FEA tension specimens ............ 163

5-11 The von Mises stress distribution in the forged FEA specimen .............. 164

5-12 Evolution of the axial stress vs. equivalent plastic strain for tension-compression cyclic loading. .................................................................... 166

5-13 Geometry of the thin-walled specimen used for free-end torsion loading with dimensions expressed in millimeters. ............................................. 168

5-14 Isocontour of the predicted axial displacement that develops during free-end torsion loading of a forged ES-1 material: ....................................... 169

5-15 Longitudinal elongation by shear strain for forged quasi-static specimens under quasi-static torsion. ...................................................................... 170

5-16 The finite element SHPB and an inset containing a close-up of an ES-1 specimen. ............................................................................................... 171

5-17 SHPB incident, transmitted, and reflected waves from forged specimen and FEA. ................................................................................................ 172

5-18 Equivalent plastic strain isocontour in the cylindrical Taylor impact specimen of forged ES-1. ....................................................................... 175

5-19 Linear relationship of deformed section length by deformed total length following experimental and simulated Taylor impact tests of specimens of ES-1. ...................................................................................................... 176

5-20 Stress strain response of ES-1 subject to dynamic strain rates in Taylor impact tests conducted by Torres et al. .................................................. 177

Page 13: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

13

LIST OF ABBREVIATIONS

AHSS Advanced high strength steel

AFRL Air Force Research Laboratory

AISI American Iron and Steel Institute

AUST SS Austenitic stainless steel

AWEF Advanced Weapons Effects Facility

BCC Body-centered cubic

BCT Body-centered tetragonal

C Celsius

CPB Cazacu, Plunkett, and Barlat

CDIC Circular polarized light differential interference

CP Complex-phase

DAS Dendritic arm spacing

DIC Digital image correlation

DP Dual-phase

EBSD Electron backscatter diffraction

EDS Energy dispersion spectrography

F Fahrenheit

FCC Face-centered cubic

FEA Finite element analysis

ft Foot

GPa Gigapascal

HIP Hot isostatic pressure

ksi Kilopounds per square inch

LVDT Linear variable differential transformer

Page 14: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

14

m Meter

MHz Megahertz

mm Millimeter

MPa Megapascal

MRD Maximum multiple random distribution

ms Millisecond

MS Martensitic steel

OFHC Oxygen-free high thermal conductivity

Pa Pascal

psi Pounds per square inch

REEF Research and Engineering Education Facility

RGB Red, green, and blue

RVE Representative volume element

s Second

SEM Scanning electron microscope

SHPB Split Hopkinson pressure bar

STF Strain-to-failure

STN Strain-to-necking

STP Standard test procedure

TRIP Transformation-induced plasticity

TWIP Twinning-induced plasticity

UF University of Florida

UHSLA Ultra high strength low alloy

UHSS Ultra high strength steel

UMAT User material subroutine

Page 15: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

15

USAF United States Air Force

V Voltage

Page 16: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

16

Abstract of Dissertation Presented to the Graduate School Of the University of Florida in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy

EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

STRENGTH MARTENSITIC STEELS BASED ON A NEW DISTORTIONAL HARDENING MODEL

By

Elizabeth K Bartlett

May 2018

Chair: Oana Cazacu Major: Mechanical Engineering In this dissertation is presented an extensive experimental and theoretical

investigation into the mechanical behavior of ultra high strength martensitic

steels, with the overall goal of determining the effect of processing on the

mechanical properties.

For this purpose, experimental characterization of forged, cast, and cast

and hot-isostatically pressed (HIP) specimens was conducted for both quasi-

static and high-rate loadings. For quasi-static loadings, the influence of loading

history was quantified by performing monotonic compression, monotonic tension,

and cyclic tension tests. Moreover, for monotonic tensile tests the influence of the

specimen geometry on localization of the deformation and strain-to-failure was

investigated. While the forged material demonstrated the greatest initial yield and

flow stress for all loading conditions, the test results indicate that the ductility of

the cast material was significantly increased by the subsequent HIP. Irrespective

of processing, the martensitic steel studied displays a higher plastic flow stress in

Page 17: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

17

uniaxial compression than in uniaxial tension, this strength differential effect

evolving with accumulated plastic deformation.

Comparison between the quasi-static and dynamic flow stresses

indicated that the dynamic increase factor is approximately 1.1 for all three

materials.

On the basis of the experimental results, an elastic/plastic modeling

approach was adopted. To account for tension-compression asymmetry in

yielding the isotropic form of Cazacu and Barlat (2004) yield criterion was used.

As concerns hardening of martensitic steels, a new distortional hardening model

was developed. In contrast to the existing hardening models, the new model

proposed can account for the Bauschinger effects.

Finally, the theoretical model was implemented in a fully three-

dimensional, implicit finite element solver, and the model predictions were

validated through comparison with data that were not used for identification of the

model parameters.

Page 18: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

18

CHAPTER 1 INTRODUCTION

This dissertation is devoted to the experimental characterization and

modeling of the effect of the manufacturing processes on the mechanical

behavior of a high strength martensitic steel, Eglin steel or ES-1. In this chapter,

an overview of the state-of-the-art on the metallurgical studies, experimental

mechanical characterization, and modeling of high strength steels is presented.

The key open research issues are identified along with the gaps in knowledge

that are attempted to be filled by the research conducted as part of this

dissertation.

1.1 Background of High Strength Martensitic Steels

Steels, alloys of iron and carbon, are some of the most widely used

engineering materials due, in part, to its strength and ease of manufacture.

Generally, the production costs of steel remain low because of an abundance of

suppliers and well-established supply lines. Although steel is viewed as nearly

an obsolete material by some who are more interested in the attractive

technology of, say, composite materials, the modern steel industry is using

progressive research and providing new alloys with improved mechanical

properties. This continuing progress in steel alloys development is due in a large

part to ever-changing safety and environmental requirements affecting the

contemporary automotive industry. A large class of new steels with yield

strength greater than 1379 MPa—known as advanced high-strength steels

(AHSS) or ultra high-strength steels (UHSS) have been developed over the last

decades. It encompasses dual-phase (DP) steels, complex-phase (CP) steels,

Page 19: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

19

transformation-induced plasticity (TRIP) steels, twinning-induced plasticity

(TWIP) steels, and martensitic steels (MS).

The evolution of UHSS can be generally considered as comprising three

distinct generations. The first-generation UHSS steels are well accepted and are

currently widely in use in myriad applications. Examples include DP, CP and

TRIP. Other steels that are currently in transition from development into

production include TWIP and austenitic stainless steel (AUST SS). These

second-generation UHSS have high strength, comparable to MS, and display

ductility similar to mild steel. Third-generation steels are currently in research

and development, the aim being to fill the gap in terms of properties which is

shown in the strain-to-failure (STF) vs. yield stress graph of Figure 1-1.

Figure 1-1. Tensile yield stress vs. strain-to-failure of current steels and expected strengths of third-generation UHSS.

Eglin steel is a third-generation martensitic steel that has been developed

for use under dynamic conditions where exceptional strength and toughness are

required (e.g. as a casing material that needs to survive the high-impact speeds

Page 20: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

20

during deep penetration events). Iron is the base metal material and can take on

the crystalline forms (allotropes in Figure 1-2): body-centered cubic (BCC) and

face-centered cubic (FCC), depending on its temperature. The BCC form of iron

is called ferrite (or -iron). Below 910oC, the BCC allotrope of pure iron is stable.

Above this temperature the face-centered cubic allotrope of iron, called austenite

(or -iron) is stable. Rapid cooling (quenching) of the austenite form of iron in oil

or water at such a high rate that carbon atoms do not have time to diffuse results

in the FCC austenite transforming into a body-centered tetragonal (BCT) form,

called martensite that is supersaturated with carbon. The shear deformations

taking place produce a large number of dislocations, which is a primary

strengthening mechanism of the resulting martensitic steels.

Figure 1-2. The unit cell of single-crystal microstructures of solid steel. A) FCC, B) BCC, and C) BCT.

Moreover, the type of crystal structure present dictates the stress-strain

response of the material and its ability to deform plastically (i.e. sustain

permanent deformation). Irreversible or plastic deformation occurs by

crystallographic slip or twinning, mechanisms illustrated schematically in

Page 21: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

21

Figure 1-3. During crystallographic slip, a lattice plane glides by an adjacent

plane along a specific direction. Slip occurs on densely packed planes, and

therefore the slip systems are different depending on the single-crystal structure.

An imperfection in the three-dimensional lattice is known as a dislocation. The

energy required for the activation of crystallographic slip is significantly reduced

in the vicinity of a dislocation. For example, in pure iron, the crystal structure has

relatively little resistance to the iron atoms slipping past one another, and so pure

iron is quite ductile, or soft and easily formed. In steel, small amounts of carbon,

other elements, and inclusions within the iron act as hardening agents that

prevent the movement of dislocations that are common in the crystal lattices of

iron atoms. Another mechanism for plastic deformation in which the lattice

structure rotates about a plane of symmetry is known as twinning (see

Figure 1-3). Cubic single-crystal structures generally require more energy to

produce twinning and so twinning is less prevalent until temperatures dip well

below ambient (<196oC) or very high strain rates (>106 /s) are produced (Blewitt

et al., 1957, Huang et al., 1996).

Page 22: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

22

Figure 1-3. Illustration of plastic deformation mechanisms. A) slip and B) twinning.

Additionally, manufacturing processes can affect the microstructure and

subsequent mechanical response of martensitic steels, and specifically ES-1.

Forging is a forming process that involves applying large compressive forces with

a power hammer, press, or die to work a piece to deform it into a predetermined

shape. Deformation of the metal is accomplished using hot, cold, or even warm

forging processes. Ultimately, the manufacturer will look at a number of criteria

before choosing which type of forging is best for a particular application. Forging

Page 23: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

23

reduces the retained porosity of the material, reduces segregation, and realigns

grains to increase the specimen strength.

Casting is a process in which a liquid metal is poured into a mold that

contains a hollow cavity of the intended shape. The metal and mold are then

cooled, and the metal part (the casting) is extracted. Casting including

a solidification process, the solidification phenomenon controls most of the

properties of the casting. Moreover, most of the casting defects occur during

solidification, such as gas porosity and solidification shrinkage. To reduce this

retained porosity, some castings are subjected to hot isostatic pressing (HIP). In

the HIP process, the casting is subjected to both elevated temperature and

isostatic gas pressure in a high-pressure containment vessel.

1.2 Survey of the Experimental Studies on Martensitic Steels

At present, most of the studies on martensitic steels report only the

mechanical response under tensile loadings (e.g. Little et al., 1978). However,

early researchers have investigated the mechanical response under both tensile

and compressive loadings, and concluded that martensitic steels have higher

flow stress in compression than in tension. A brief summary of the main findings

is presented in the following.

Leslie and Sober (1967) conducted experiments on a dozen different

materials including SAE 4300 steels under uniaxial tension and compression, at

temperatures between -195oC and 23oC, and strain rates between 10-5 and 10-1

s-1. It was found that untempered martensites are more resistant to plastic flow in

compression than in tension during initial yielding and the early stages of plastic

Page 24: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

24

flow. To quantify this tension-compression asymmetry in the plastic regime,

these authors introduced a parameter, called the fractional strength differential

parameter (FSD), which is defined as:

C T

C T

FSD

, (1-1)

where σC denotes the flow stress in uniaxial compression and σT is the flow

stress in uniaxial tension. It was observed that for SAE 4300 series 4310, 4320,

4330, and 4340 steels there is a linear relationship between the fractional

strength differential parameter and the carbon content by percent weight (see

also Figure 1-4).

Figure 1-4. The fractional strength differential parameter vs. the carbon content by percent weight for different SAE 4300 series.

Several explanations for the observed strength differential in martensitic

steels have been proposed. One of the explanations is that SD effects are due

Page 25: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

25

to solute-dislocation interaction, namely that the resulting martensitic BCT

structure, specifically the ratio c/a of the cell (see Figure 1-2) leads to non-linear

local elastic strains. Using a conservative estimate of the carbon–dislocation

interaction energy, Hirth and Cohen (1970) demonstrated that the resulting SD

values are between 3 and 6%, a range similar to that observed experimentally.

Chait (1971) conducted uniaxial tension and compression tests of 4340

and H11 high-strength steels and 410 martensitic stainless steels at room

temperature. A constant value of the SD parameter (see Eq. (1-1)) was

observed over a large range of plastic strains for all steels except the stainless

steel. Furthermore, Chait (1971) reported that for SAE 4340 the SD parameter

C TSD decreases with increasing tempering temperatures as shown in

Figure 1-5.

Figure 1-5. The effect of tempering temperature on the strength differential of quenched and tempered 4340 UHSS.

Page 26: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

26

Resistance to plastic strain increases with increasing strain rates. For

example, Nadai (1963) reported that for 0.35 carbon steel there is a two-fold

increase in flow stress as the strain rate was increased 10,000 times. More

comprehensive surveys on the strain-rate sensitivity of various materials,

including high-strength steels were conducted by Maiden and Green (1966), Last

et al. (1996), and Johnson and Cook (1983).

Specifically, Maiden and Green (1966) conducted an investigation of the

rate-sensitivity of several metallic and non-metallic materials using a medium

strain-rate machine and a split Hopkinson pressure bar (SHPB) (for a description

of this dynamic testing technique and apparatus, see Chapter 3) over a range of

strain rates 10-3 to 104 s-1. They found that aluminum alloys 7075-T6 and 6061-

T6 were rate insensitive within this range. The rate-dependent behavior of 6A1-

4V titanium was described using a power law type relation between the uniaxial

compressive flow stress normalized by the quasi-static compressive yield stress

and the strain rate, , i.e.

1 n

C

k

. (1-2)

In the above equation, k and n are constants. For 6A1-4V titanium, k=1

and n = 0.2.

Johnson and Cook (1983) presented a new equation for description of the

effects of temperature and strain rate on the flow stress of materials based on

data collected in quasi-static uniaxial tension tests and SHPB tensile tests at

various temperatures. The materials tested included oxygen-free high thermal

conductivity (OFHC) copper, Cartridge brass, Nickel 200, Armco tool, Tungsten

Page 27: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

27

alloy, and DU-.71Ti depleted uranium. This empirical relationship was validated

using Taylor impact tests in which strain rates are typically between 104 and 105

s-1. The isothermal form of Johnson-Cook equation is:

*1 lnn

J J P J PA B C (1-3)

In the above equation, AJ, BJ, CJ, and n are material parameters, and *

p is

the relative plastic strain rate defined as:

*

0

p

p

p

, (1-4)

Where p is the dynamic plastic strain rate while 0p is the quasi-static plastic

strain rate.

Last et al. (1996) characterized the high strain-rate response of Hy100,

Hy130, and AF-1410, all martensitic steels. Hy100 was rate sensitive regardless

of temperature. However, Last et al. (1996) found that the rate-sensitivity of

HY130 and AF-1410 increased with temperature as shown in Figure 1-6.

Page 28: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

28

Figure 1-6. The effect of tempering on the flow stress of various steels.

Understanding and modeling the behavior of engineering materials, and in

particular steels is a key for estimating the service life and safety of structural

components.

The stress-strain response of most engineering metals under cyclic

loading is often significantly different from the monotonic response. In fact, the

stress-strain response observed during the first cycle of loading is often

significantly different from the second, and the second is different from the third,

until the material stabilizes. In the subsequent hysteresis loops of Ti-8A1-1Mo-

1V (Collins, 1981), the difference between the stress-strain response in the 100th

and the 1000th cycle is less pronounced than that of the first and third.

Eventually, the same flow stress is observed in subsequent cycles (i.e. the

material has stabilized) and a stable hysteresis loop is observed. Thus, the

Page 29: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

29

stress evolves under subsequent reversals until it reaches stability as

demonstrated by the stable hysteresis loop.

For martensitic steel SAE 4340, Collins (1981) reported stable hysteresis

loops under completely reversed strain. Initially the material was subject to a

tensile strain beyond yielding until the desired strain amplitude was attained.

Then, the displacement was reversed, unloading and reloading in compression

within the elastic regime to the flow stress in compression, and then to the same

strain amplitude, but this time in compression. Next, the displacement is

reversed again, this time back to tension. The ability of a material to resist

repeated straining is characterized through a cyclic stress-strain curve, loci of the

maximum values of stress and strain from stable hysteresis loops at different

levels of strain amplitude.

For SAE 4340 the monotonic flow stress is greater than the cyclic stress,

meaning that SAE 4340 softens under cyclic loading by up to 30%.

Landgraf (1970) conducted a comprehensive experimental study in which

he compared the hardening behavior of various metals, including martensitic

steels, under monotonic and cyclic loadings. Moreover, for any given material he

identified a power law type hardening model and reported the values of the

respective coefficients. In Figure 1-1 are presented these results. Note that the

high-strength materials (including martensitic steels) typically soften under cyclic

loading. In addition, Landgraf (1970) also reported the monotonic and cyclic

stress-strain behavior.

Page 30: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

30

Table 1-1. Monotonic and cyclic material properties. Material Condition Monotonic

Yield Strength (MPa)

Cyclic Yield

Strength (MPa)

Monotonic Strain

Hardening Exponent

Cyclic Strain

Hardening Exponent

Cyclic Behavior

OFHC annealed 21 138 0.400 0.150 Hardens

partial annealed 255 200 0.130 0.160 Stable

cold worked 345 234 0.100 0.120 Softens

2024 Al T4 303 448 0.200 0.110 Hardens

7075 AL T6 469 517 0.110 0.110 Hardens

Man-Ten steel as-received 379 345 0.150 0.160 Both

SAE 4340 350 BHN* 1172 758 0.066 0.140 Softens

Ti-8Al-1Mo-1V duplex annealed 1000 793 0.078 0.140 Both

Waspaloy Ref 11 545 703 0.110 0.170 Hardens

SAE 1045 595 BHN 1862 1724 0.071 0.140 Stable

500 BHN 1689 1276 0.047 0.012 Softens

450 BHN 1517 965 0.041 0.150 Softens

390 BHN 1276 758 0.044 0.170 Softens

SAE 4142 670 BHN** 1620 N/A 0.140 N/A Hardens

560 BHN 1689 1724 0.092 0.130 Stable

475 BHN 1724 1344 0.048 0.120 Softens

450 BHN 1586 1069 0.040 0.170 Softens

380 BHN 1379 827 0.051 0.180 Softens

*Quenched and tempered **As-quenched

The Bauschinger effect, illustrated in hysteresis loops, is so named after

Bauschinger (1886) who observed that “after the stress is reversed from tension

to compression both the elastic range and a yield point for the reversed direction

of straining have completely disappeared.” In other words, a material subject to

uniaxial tension will harden in tension, but upon load reversal, the flow stress in

compression is decreased. If the material is loaded in uniaxial compression, it

hardens in compression and the tensile flow stress is decreased. The

Bauschinger effect is also seen in specimens that have been pre-strained. As an

example, Paul (1968) presented the stress-strain response of copper in uniaxial

tension tests and subsequent compression tests at room temperature from tests

by Nadai (1963).

Page 31: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

31

To study the Bauschinger effect, the limits of the elastic region under

uniaxial tension and compression were calculated based on the stable hysteresis

loops of SAE 4340 presented by Collins (1981) using 0.1% offset following load

reversal or the maximum absolute flow stress prior to load reversal. Figure 1-7

plots the absolute flow stress in uniaxial tension and uniaxial compression under

completely reversed strain for SAE 4340. Notice the Bauschinger effect in Figure

1-7 for SAE 4340.

Figure 1-7. Absolute flow stress vs. plastic strain of SAE 4340 from hysteresis loops of cyclic testing.

Bridgman (1952) collected extensive experimental data at high hydrostatic

pressure and concluded that the yield stress of metals is unaffected by

hydrostatic pressures up to 2.8 GPa. However, Bridgman demonstrated

immense increases in ductility under high hydrostatic pressure. For example, he

Page 32: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

32

estimated true strain of 4.4 106 in the neck of steel tension specimens under

hydrostatic pressure of 9.8 GPa.

With an understanding of the microstructure and mechanical behavior of

martensitic steels, a brief history and survey of Eglin steel, specifically, follows in

Section 1.3.

1.3 Eglin Steel, ES-1

AF-1410 is a high-strength, high-toughness, super alloy steel for use in

aerospace applications. In an effort to find a lower-cost substitute for AF-140,

Ellwood National Forge Company developed a new class of high-strength, low-

alloy martensitic steel called Eglin steel (ES-1). ES-1 received the patent

US7537727 Eglin Steel—A Low Alloy High Strength Composition by Morris

Dilmore and James Ruhlman (2009).

Table 1-2 presents the chemical composition of a high-alloy steel AF-

1410, UHSLA standard SAE 4340, and that of ES-1 for comparison.

Page 33: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

33

Table 1-2. Chemical composition by percent weight of AF-1410, SAE 4340, and ES-1.

Element AF-1410* (%)

SAE 4340** (%)

ES-1 (%)

Cobalt 13.5-14.5000 0.000 0.000 Nickel 9.5-10.5000 1.65-2.000 5.000 Chromium 1.8-2.2000 0.7-0.900 2.380 Molybdenum 0.9-1.1000 0.2-0.300 0.550 Carbon 0.13-0.1700 0.37-0.430 0.250 Manganese <0.1000 0.6-0.800 0.850 Silicon <0.1000 0.15-0.300 1.250 Aluminum 0.0000 0.000 0.000 Vanadium 0.0000 0.000 0.180 Tin <0.0020 0.000 0.000 Lead <0.0020 0.000 0.000 Zirconium <0.0020 0.000 0.000 Boron <0.0005 0.000 0.000 Sulphur <0.0050 <0.040 0.012 Phosphorus <0.0080 <0.035 0.015 Tungsten 0.0000 0.000 1.980 Copper <0.1000 0.000 0.500 Calcium 0.0000 0.000 0.020 Nitrogen <0.0015 1.65-2.000 0.140

Iron Bal. Bal. Bal.

*AF-1410 chemical composition from Little (1979). **SAE 4340 chemical composition from Lynch (2011).

Notice the total alloy content of AF-1410, SAE 4340, and ES-1 are as high

as 29, 6, and 13%, respectively. Also, note that ES-1 uses roughly half as

much nickel and cobalt as other superalloys (e.g. AF-1410), which makes it a

low-cost alternative. Although the percentage of nickel was decreased in ES-1

compared to AF-1410, substituting silicon to help with toughness and particles

of vanadium carbide and tungsten carbide for additional hardness and high-

temperature strength ensured that ES-1 has the required properties. Moreover,

ES-1 contains less carbon than SAE 4340, and thus the tetragonality of ES-1

(ratio of c/a of its BCT structure, see also Figure 1-2) is just 1.01.

In addition to smaller amounts of expensive alloys, Eglin steel’s low cost is

in part a result of its manufacturability and machinability. Ingots of ES-1 are

manufactured by several processes depending on the application requirements;

Page 34: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

34

these processes include electric arc, ladle refined, vacuum treated, vacuum

induction melting, vacuum arc re-melting, and/or electro-slag re-melting.

However, the ingot manufacturing process does affect the mechanical properties.

Final Eglin steel products can be produced by forging, casting, extrusion, rolling

or other conventional methods in accordance with the patent.

Before publishing the patent, Eglin steel was tested in quasi-static and

dynamic tension at ambient and high temperatures. The Rockwell hardness and

Charpy V-notch energy were also measured. Table 1-3 presents the ultimate

tensile strength, initial yield stress, and strain-to-failure in uniaxial tension of ES-1

along with the same properties of other martensitic steels: AF-1410 and SAE

4340.

Table 1-3. Mechanical properties under uniaxial, quasi-static tensile loading of AF-1410, SAE 4340, and ES-1.

Ultimate Tensile Strength

(%)

Yield Strength (%)

Strain to Failure (%)

AF-1410 1620 1482 12.0 SAE 4340 1793 1496 10.0 ES-1 1701 1337 18.4

* AF-1410 mechanical properties from Little (1979) ** SAE 4340 mechanical properties from Lynch (2011).

Note that the largest percentage difference in both ultimate tensile and

yield strength between the three UHSS is approximately 10%. The reported

strain-to-failure of ES-1 is significantly higher than AF-1410 and SAE 4340. SAE

4340 has the highest initial yield strength and ultimate tensile strength.

The graph in Figure 1-8 shows the tensile yield stress vs. STF of several UHSS

including ES-1. Notice that Eglin steel is a high strength steel with twice the

ductility of other martensitic steels.

Page 35: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

35

Figure 1-8. STF vs. tensile yield stress for different steels including conventional steels and UHSS.

Boyce and Dilmore (2009) used a servohydraulic motor outfitted with a

custom-built load cell to characterize the stress-strain response of flat specimens

of AerMet100, 4340M, HP9-4-20 and ES-1c in both the quasi-static and dynamic

regimes. They used strain gages in a custom built load cell on the upper and

lower threaded sections of the test specimens to measure the load in addition to

the standard quartz load cell that introduces oscillations in the force data.

Secondly, a rubber pulse shaper was used in the high strain-rate experiments (1

to 200 s-1) to reduce the elastic oscillations in the test. Using both of these

methods the oscillations were dampened enough to directly determine the 0.2%

offset from the stress-strain curve. All of the alloys exhibited slight strain-rate

sensitivity (magnification factors between 1.06 – 1.10 at 2% strain). The data

was used to obtain the coefficients of a power law hardening for strain-rate

sensitivity (see Eq. (1-2)), and a semi-logarithmic relationship, which is more

typical for ferrous materials. It is worth noting that among all the alloys on which

Page 36: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

36

the authors reported data, Eglin steel was the only alloy that shows increase in

ductility with increasing the strain rate.

In addition to the experimental characterization completed by Dilmore and

Ruhlman (2009) and included in the patent, several other researchers conducted

dynamic characterization experiments on ES-1. Importantly, Torres et al (2009)

completed twenty Taylor impact tests with 164-caliber (41.66 mm) and 215-

caliber (54.61) cylindrical specimens to determine the stress-strain response of

Eglin steel at strain rates between 103 and 104 s-1 which are higher than those

that can be achieved with the SHPB apparatus, but lower strain rates that occur

in plate impact tests. To interpret the data, use was made of the one-

dimensional theory proposed by Jones et al. (1998). However, it was found that

a 1-D analysis is no longer accurate for strains larger than 10%. Nevertheless, it

was possible to estimate the dynamic flow stress at 10% deformation, which

appears to be approximately 2.5 GPa.

Martin et al. (2012) conducted five symmetric plate impact tests, with the

impact velocities being between 400 and 1000 m/s using ES-1 impactors and

targets. Prior to conducting the tests, Martin et al. (2012) used ultrasonic

techniques to determine the average longitudinal and shear wave speeds of the

material. It was found that ES-1 has average longitudinal and shear wave speeds

of 5.88 and 3.20 km/s and bulk wave speed of 4.58 km/s (Poisson ratio 0.29).

The impactors were launched from a smooth bore powder gun and a velocity

interferometer (VISAR) was used to measure the velocity of particles on the back

side of the targets as shown in Figure 1-9.

Page 37: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

37

Figure 1-9. ES-1 symmetric plate impact experimental setup of Martin et al.

The Hugoniot elastic limit σHEL is defined as:

0

2

L EHEL

C u

(1-5)

where ρ0 is the initial density, CL is the longitudinal wave speed, and uE is the

particle velocity where the elastic to plastic transition occurs. For ES-1, σ HEL is

approximately 2.35 GPa.

The spall strength (dynamic tensile strength) is defined as:

0

2

L pbC uS

(1-6)

where Δupb denotes the pullback signal.

Estimating the pullback signal as being Δupb~0.279 km/s, the spall

strength of ES-1 calculated with the above formula is: S~6.34 GPa.

Finally, in experiments conducted by Weiderhold, Lambert, and Hopson

(2010), the detonation of hollow explosive-laden cylinders of Eglin steel was used

to analyze fragmentation and develop a statistical failure model.

In summary, prior to the research done as part of this dissertation the

stress-strain response has been experimentally characterized only under uniaxial

loading. In reality, materials are subjected to complex three-dimensional loads.

Page 38: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

38

As materials cannot be tested in every possible loading configuration, a model is

required to predict behavior under complex, three-dimensional loads. Ultra high-

strength martensitic steels exhibit a well-defined yield stress and plastic

deformation. Therefore, the most appropriate framework for modeling their

behavior is that of the elastic-plastic theory. The general form of the governing

equations is briefly presented in the following, along with the state-of-the art in

description of yield criteria and hardening laws for metallic materials.

1.4 Elastic-Plastic Modeling

1.4.1 Yield Criteria

Elastic-plastic models best describe polycrystalline materials that have a

well-defined yield stress that defines the onset of plastic deformation. Yield

criteria define the boundary between the elastic and plastic regimes, and can be

visualized in six-dimensional stress space by surfaces. In the interior of the yield

surface the material behavior is elastic, usually characterized by Hooke’s linear

relationship. On the yield surface, the material behavior is plastic, nonlinear, and

depends on internal variables such as loading history. The yield surface grows,

translates, and changes shape based on the hardening behavior of the material.

Therefore, the yield function (F) is commonly expressed as the difference

between the effective stress ( ), a function of the Cauchy stress and a measure

of the equivalent plastic strain, and the hardening (Y), which is a function of only

the equivalent plastic strain:

, , 0p p pF Y σ σ (1-7)

Page 39: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

39

The earliest yield criterion for isotropic pressure-insensitive metals was

proposed by Henri Tresca in 1865. It states that yielding occurs when the

maximum shear stress reaches a critical value, f . This criterion can be

expressed in terms of the principal stresses as:

1 2 2 3 1 3max( , , ) 2 f . (1-8)

The Tresca yield criterion is independent of the hydrostatic pressure (or

mean stress) and therefore can also be written in terms of the deviatoric stress

and the surface can be presented in the three-dimensional principal stress space

(or Haigh-Westergaard space) as an infinite, hexagonal prism with the

longitudinal axis (Figure 1-10) coinciding with the hydrostatic axis, or as a regular

hexagon in the deviatoric or plane, (plane having the normal the hydrostatic

axis).

1 2 3, ,

Page 40: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

40

Figure 1-10. Tresca yield surface in Haigh-Westergaard space with the hydrostatic axis and deviatoric plane.

Richard Edler von Mises proposed in 1913 another yield criterion for

isotropic materials which states that yielding occurs when J2, the second invaraint

of the stress deviators reaches a critical value. It follows that the criterion writes:

2 2 2 2

2 1 2 3

1/ 3

2TJ s s s (1-9)

where s1, s2, s3, are the principal values of the stress deviator s and T is the

yield stress in uniaxial tension. In the Haigh-Westergaard space, the von mises

yield surface is a right circular cylinder of generator the hydrostatic axis (Figure

1-11).

Page 41: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

41

Figure 1-11. Von Mises yield surface in Haigh-Westergaard space with the longitudinal axis as the hydrostatic axis.

In 1949, Daniel C. Drucker proposed a yield criterion for isotropic

materials that involves both J2 and the third-invariant 3 1 2 3J s s s of the deviatoric

stress. It is expressed as:

3 2 6

2 3 fJ cJ (1-10)

where c is a material parameter.

The projections in the biaxial plane ( 3 0 ) of the yield surfaces of the

Drucker (1949) corresponding to c=2.25, Mises, and Tresca yield criteria in the

plane of the third principal stress are presented in Figure 1-12.

Page 42: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

42

Figure 1-12. Projection in the biaxial plane of the yield surfaces of Drucker, Mises, and Tresca yield criteria.

Note that according to either criterion the yield stress in tension is equal to

the yield stress in compression.

In 1954, Hershey proposed an yield function later used by Hosford (1972)

expressed in terms of principal stresses.

The criterion involves a unique parameter n. It reduces to Tresca's

criterion for n=infinity and to von Mises for n=2.

1

1 2 2 3 1 3

1 1 1

2 2 2

nn n n

(1-11)

In 2004, Cazacu and Barlat formulated an isotropic yield criterion

represented by an odd function that depends on both invariants J2 and J3. In this

manner, it was possible for the first time, to account for tension-compression

asymmetry in yielding of pressure-insensitive metals. Its expression is:

Page 43: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

43

13 322 33

3 3 2T

J J

c

. (1-12)

The coefficient, c, can be determined based on the flow stress in tension

( ) and compression ( ) as (see Cazacu and Barlat, 2004):

3 3

3 3

3 3

2

T C

T C

c

. (1-13)

The Cazacu and Barlat (2004) isotropic yield function reduces to von

Mises for . For , the Cazacu and Barlat (2004) yield criterion is a

triangular prism with a longitudinal axis coinciding with the hydrostatic axis

(Figure 1-13).

T C

0c 0c

Page 44: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

44

Figure 1-13. Cazacu and Barlat yield surface in Haigh-Westergaard space.

The projections in the biaxial plane ( 3 0 ) of the Cazacu and Barlat

(2004) yield surface is presented in Figure 1-14 for various ratios of flow stress

( /T C =3/4, 1, and 5/4) which correspond to c=1.056, 0.000, and 0.082,

respectively.

Page 45: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

45

Figure 1-14. Projection in the biaxial plane of the Cazacu and Barlat yield surface corresponding to flow stress ratios of 3/4, 1, and 5/4.

Porosity can produce pressure dependence on yielding. Gurson deduced

the yield criterion for a von Mises material containing randomly distributed pores.

The pores were either of spherical or cylindrical geometry. In 2009, Cazacu and

Stewart (2009) developed a yield criterion for isotropic materials that display

tension-compression asymmetry based on Cazacu, Plunkett, and Barlat (2006),

also known as CBP-06, rather than von Mises effective stress:

2

2

2

2

32 cosh 1 0

2

32 cosh 1 0

2

e mm

T T

e mm

T C

f f

f f

(1-14)

where

is the mean macroscopic stress, m

Page 46: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

46

is the effective macroscopic stress,

is the absolute value of the matrix material yield stress under uniaxial

compression,

is the matrix material yield stress in uniaxial tension, and

f is the void volume fraction.

This is not an exhaustive list of yield criteria, and many other criteria exist

to model materials with different characteristics. For example, several models for

nonmetallic materials include the first invariant of the stress to account for

pressure sensitivity. Furthermore, many materials display orthotropic behavior

and are better described by orthotropic criteria such as that of Hill (1948) and

Cazacu, Barlat and Plunkett (2006).

In this research, a new elastic-plastic model based on hardening as a

function of the second and third invariants of plastic strain using Cazacu and

Barlat (2004) asymmetric yield criteria is developed with a completely new

approach to the definition of asymmetry. Therefore, a background of existing

hardening models is presented in the subsequent section.

1.4.2 Background of Hardening Models

The term hardening often evokes the uniaxial stress-strain response of

materials under monotonic, uniaxial tension. Under this loading, a simple power

law approach can be used to describe the hardening or the increase in flow

stress following initial yielding of the material. In three dimensions, hardening is

characterized by subsequent yield surfaces that satisfy the yield function (Eq

(1-15)). Paul (1968) describes the three-dimensional hardening as isotropic if the

e

C

T

Page 47: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

47

subsequent yield surfaces are larger than the previous yield surface without a

change in shape or translation in stress space and kinematic if the subsequent

yield surfaces translate in Haigh-Westergaard space. In 1957, Hodge proposed

combined hardening with an isotropic component (Y) and a kinematic portion

resulting from a fictitious back stress ( ):

, , ,p p pF Y σ σ (1-15)

1.4.2.1 Isotropic hardening

The isotropic hardening, defined as the expansion of the yield surface of

metallic materials under monotonic loading is generally considered to be a

function of equivalent plastic strain and can be described utilizing the hardening

curve of the material observed in uniaxial tension. The equivalent plastic strain

p associated to a given yield criterion is defined by the work equivalence

principle (Hill, 1987). For example, if a material is governed by the von Mises

yield criterion, we have

2

3

p p pε : ε

. (1-16)

Typical rate-independent isotropic hardening laws are Swift (1952) law

and Voce (1948) law.

snp

s pY k (1-17)

V pCp

V VY A B e

(1-18)

In the above equations, ks and ns and AV, BV, and CV are the material

parameters.

Page 48: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

48

1.4.2.2 Kinematic hardening

To account for the Bauschinger effect, kinematic hardening under cyclic

loading is usually modeled by a translation of the yield surface resulting from the

application of a fictitious back stress tensor (). It should be noted that a

nonzero back stress destroys the symmetry based on the principal deviatoric

stresses. Several kinematic hardening models have been proposed including

Melan (1938) based on the plastic strain tensor, Prager (1955) based on the

plastic strain-rate tensor, and Lubliner (1990) with a term dependent on the

equivalent plastic strain rate presented below with material constants c and a:

p

c

c

c a

p

p

p

ε

ε

ε ρ

(1-19)

Dafalias and Popov (1975) and Krieg (1975) defined two-surface models

in which the yield surface is limited by an outer bounding surface in stress space.

1.5 Goals of Current Research

The overall goal of the dissertation research is to determine, for the first

time, the effect of three different manufacturing processes on the mechanical

response of Eglin steel: Chapter 2 presents the microstructure and porosity of the

material under study in its initial condition. Chapter 3 presents an in-depth

experimental characterization of the material response under various loading

paths and strain rates. Chapter 4 includes a review of finite element analysis, a

development of the theoretical model, identification of material parameters, and

integration of the theoretical model into a fully three dimensional finite element

analysis. Chapter 5 explains the model validation using independent

Page 49: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

49

experimental data and finite-element analyses as well as predictions of the

material response under more complex loadings and higher strain rates. Finally,

Chapter 6 concludes this dissertation and research with a brief summary and

recommendations for continuing this exciting research.

Page 50: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

50

CHAPTER 2 MICROSCOPY

Each manufacturing processing procedure implies different consequences

for the material: forging is a costly process but reduces porosity, shrinkages and

cavities, while casting is a cheaper process, but voids and cavities are created

during the solidification of the material. Most of the defects induced by

solidification could be removed with an HIP operation. Therefore, the initial

condition of each material including chemical composition, porosity, prior

austenite grain size, and dendrite arm spacing (DAS) was characterized using

microscopy.

The cast Eglin steel was received at AFRL in 2-in thick plates. The HIP

process involved heating the cast specimens to 2125oF at 15 ksi. All the of cast

specimens were then subject to homogenization to decrease segregation,

austenization to break down iron carbides, quenched to form martensite, and

tempered to relieve residual stresses. The forged specimens were subject to a

milder heat treatment process due to the superior homogeneity of forged

materials as compared to cast materials. The heat treatment schedules for all

three manufacturing processes are summarized in Table 2-1.

Table 2-1. Heat treatment schedule for ES-1. Forged Cast and HIP’d Cast

Hydrogen Bake 315oC/12 hr

HIP @ 103 MPa - 1163oC/4hr -

Homogenization 1163oC/4 hr - 1163

oC/4 hr

Sub-critical Anneal 677oC/4 hr (Air)

Austenization 1010oC/1.5 hr (Vacuum)

Quench Fast-Quench (Oil) Temper 204

oC/4 hr (Recirculating air)

Page 51: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

51

Next, four different ASTM standard test specimens were machined from

the ES-1 plates: 25.4-mm flat tension, 25.4-mm round tension, and round

compression with diameters of 5.08 mm and 7.62 mm. An ES-1 specimen of

each manufacturing process was sectioned, mounted in polyfast resin using a

Struers ProntoPress-22, ground to plane, and polished on a Struers autopolisher

in accordance with ASM Metallurgy and Metallography Handbook as

summarized in Table 2-2. Finally, the specimens were polished in a Struers

vibratory polisher with OP-U Silica Carbide suspension at 40% for 2 hours.

Table 2-2. ES-1 grinding and polishing schedule. Process Grit Speed

(rpm) Force

(N) Time (min)

Lubricant Surface

Grind 220 300 20 5 Water N/A Grind 320 300 20 5 Water N/A Grind 500 300 20 5 Water N/A Grind 1200 300 20 5 Water N/A Polish 9 μm

3 μm 150 150

20 20

4 4

Blue Blue

MD-Largo, MD-Dur Polish

2.1 Optical Microscopy ES-1

The best way to characterize porosity in ES-1 according to Foley (2016) is

optical microscopy of polished specimens using brightfield illumination.

Therefore, each specimen of ES-1 was examined under a Zeiss Axio Observer

Z1 optical microscope to evaluate the porosity. The cast specimens contained

several large pores resulting from partially fused dendrites that may have

resulted from inadequate feeding as volume shrinkage occurred during

solidification. The relatively large pores in the cast material were examined using

Circular polarized light differential interference contrast (CDIC) to emphasize the

depth of the pores. Several examples are displayed in Figure 2-1 and the entire

surface is shown in Figure 2-2.

Page 52: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

52

Figure 2-1. Z-stack of several pores on the polished surface of cast eglin steel specimen using CDIC.

Page 53: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

53

Figure 2-2. Photomontage stitched from nine individual images of the polished surface of cast ES-1 specimen containing several pores.

An image processing program, Image J, was used to count (Np) and

measure the average area (Ap) of pores above 1.175 μm2 in a 2.5 mm2 area of

the materials at 200 times magnification as shown in Figure 2-3. Assuming a

spherical representative volume element (RVE), the two dimensional measured

area volume fraction of the polished surface was used to calculate the RVE inner

Page 54: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

54

(a) and outer (b) radii and the void volume fraction ( ). As expected the

void volume fraction of the forged material is insignificant while that of the cast

and cast and HIPd materials is 0.007 and 0.001, respectively.

Figure 2-3. Polished surface of ES-1 materials at 200 times magnification in an optical microscope with brightfield illumination. A) forged, B) cast and HIPd, and C) cast.

To measure the dendritic arm spacing (DAS), a sample of the cast

material was etched with Nitol and placed in a Keyence optical microscope using

brightfield illumination. Due to the tempering, the dendrites are more difficult to

identify, however, the DAS was approximated as 173 μm using the two dendrites

highlighted in Figure 2-4. The dendrites of the cast and HIPd and forged

specimens were severely deformed in the high pressure operations and thus the

DAS was not measured.

3

3v

af

b

Page 55: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

55

Figure 2-4. Nitol etched cast material surface at 100 times magnification in Keyence optical microscope under brightfield illumination.

2.2 Material Characterization using SEM

The microscopy specimens were then mounted in an FEI Quanta 200F

scanning electron microscope (SEM) and subject to a vacuum for energy

dispersion spectrography (EDS) to determine the approximate chemical

composition of ES-1 developed by each manufacturing process shown in Table

2-3.

Page 56: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

56

Table 2-3. Chemical composition of ES-1 for EDS.

Elements Eglin Steel* (%)

Forged Cast Cast and HIP’d

Nickel <5.000 1.00 0.98 1.07 Chromium 1.5-3.250 2.57 2.56 2.70 Molybdenum <0.550 0.60 0.67 0.69 Carbon 0.16-0.350 2.87 1.83 1.52 Manganese <0.850 1.06 0.50 0.58 Silicon <1.250 1.20 1.11 1.22 Vanadium 0.05-0.300 0.14 0.08 0.08 Sulphur <0.012 0.00 0.00 0.00 Phosphorus <0.015 0.00 0.00 0.00 Tungsten 1.17-3.250 1.00 0.87 0.75 Copper <0.500 0.00 0.00 0.00 Calcium <0.020 0.00 0.00 0.00 Nitrogen <0.140 0.00 0.00 0.00 Aluminum <0.050 0.00 0.00 0.00 Iron Bal. 89.55 91.41 91.39

*Nominal ES-1 composition from Boyce and Dilmore (2009).

The EDS indicates the molybdenum is slightly high and the tungsten low.

Errors in the light elements like Carbon are expected using EDS with various

elements. Then, the electron backscatter diffraction (EBSD) was conducted

within the SEM equipped with EDAX Hikari cameras to establish average grain

size, grain orientation, and texture.

The orientation maps in Figures 2-6 through 2-8 show the orientation of

BCC single-crystal grains with respect to the surface normal (z) using a red,

green, and blue (RGB) color scheme coupled to Miller indices as indicated in

Figure 2-5.

Page 57: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

57

Figure 2-5. RGB color scheme coupled with Miller indices (001), (101), and (111) in cylindrical coordinates of the specimen cross section.

Page 58: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

58

Figure 2-6. Orientation map of forged Eglin steel demonstrating prior austenite grain boundaries and martensitic lath structure.

Page 59: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

59

Figure 2-7. Orientation map of cast and HIPd Eglin steel demonstrating prior austenite grain boundaries and martensitic lath structure.

Page 60: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

60

Figure 2-8. Orientation map of cast Eglin steel demonstrating prior austenite grain boundaries and martensitic lath structure.

The prior austenite grain sizes (G) 5.0 was determined using the linear

intercept method and the circular intercept method outlined in ASTM E112-13

Standard Test Methods for Determining Average Grain Size for all three

materials.

Page 61: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

61

Additionally, EBSD was used to create pole figures and analyze the

texture of each material. Texture is the preferential alignment of grains within a

material. Texture commonly develops from material processing, but can be

mitigated by tempering. The pole figures, one figure from a plane from each

family in the cubic crystalline structure (001), (101), and (111), are presented in

Figure 2-9. The BCC orientation of all three crystal planes for all three material

processes appears to be random, confirming the samples are isotropic. The

maximum multiple random distribution (MRD) of 4.917 in the forged specimen

indicates that the (111) plane is approximately 5 times more likely to be oriented

at approximately 75o polar angle and 200o azimuth.

After the initial condition of the materials was determined, an extensive

suite of experiments was conducted to characterize the mechanical behavior of

the materials. The experimental characterization is presented in Chapter 3.

Page 62: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

62

Forged

Cast and HIP’d

Cast

Figure 2-9. Pole figures of (001), (101), and (111) for forged, cast and HIP’d, and cast Eglin steel.

Page 63: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

63

CHAPTER 3 MECHANICAL CHARACTERIZATION OF ES-1

In this chapter are presented the investigation conducted into the

mechanical behavior of ES-1 for each condition (i.e. forged, cast, cast and

HIP’d). A suite of experiments was conducted on all three ES-1 materials to

characterize the room-temperature mechanical response for both quasi-static

and dynamic strain rates, under a variety of test conditions (e.g. quasi-static and

monotonic loadings and specimen geometries). The test matrix is given in

Table 3-1.

Table 3-1. Eglin steel material characterization test matrix. Strain Rate Manufacturing

Process Loading

Configuration Specimen Shape/

Pulse Shaper Number of Specimens

Quasi-Static

Forged Tension Round 4 Pin Loaded 2

Compression Round 4 Cast and HIPd Tension Round 4

Pin Loaded 2 Compression Round 4

Cast Tension Round 4 Pin Loaded 3

Compression Round 4 High-Strain Rate

Forged Compression None 3 Pulse Shaper 2

Kinematic Forged Cyclic Round 1 Cast Cyclic Round 1 Cast and HIPd Cyclic Round 1

3.1 Hardness of Eglin Steel

Following the microstructure characterization of the initial condition of the

ES-1 materials (see Chapter 2), a Buehler Digital Hardness Tester MMT-3 with a

pyramidal indenter in Figure 3-1 was used to measure the Vickers hardness of

each sample at the AFRL Dynamic Properties Laboratory, Advanced Weapons

Effects Facility (AWEF).

Page 64: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

64

Figure 3-1. Buehler Digital Hardness Tester MMT-3.

The material resistance to plastic deformation, HV, was determined with

the following equation (Smith and Sandland, 1922)::

2

FHV

d , (3-1)

where F is the applied force and d is the average length of the diagonal of the

material indent.

Each sample was tested three times across the cross section of the

specimen at distances greater than 2.5 times the diagonal of the indentation from

the outer surface and previous indentations. The forged, cast, and cast and

HIP’d specimens had an average Vickers hardness of 529.2, 499.5, 501.7

kgf/mm2 (or 5189.68, 4920.00, and 4898.42 MPa), respectively. The Vickers

Page 65: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

65

hardness measurements converted to Rockwell Hardness are approximately

51.1, 49.1, and 49.1.

3.2 Quasi-Static Mechanical Characterization

An Instron Model 1332 mechanical test frame and an Instron Model 3156-

115 load cell at the AWEF presented in Figure 3-2 were used for both quasi-

static tension and compression testing.

Figure 3-2. AWEF quasi-static experimental characterization setup A) Instron test frame with load cell, B) pin-connected tension specimen prepared for loading and C) compression specimen prepared for loading.

Prior to each quasi-static test, the test specimen was cleaned with

acetone to remove oils and debris and painted in a high contrast, stochastic

pattern using Rustoleum black and white flat enamel spray paints to create

unique facets on the surface of the entire gage length of the test specimens.

A

B

C

Page 66: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

66

Two Point Gray high-speed cameras with 5 megapixel sensors and 75mm fixed

focal length Fuginon lenses, stationary fiber optic illuminators, and a digital delay

generator were used to record synchronized images every four seconds.

Analysis of the recorded images using digital image correlation (DIC) commercial

software, ARAMIS v6.2, reveals the location of each facet in three dimensional

space and the local, full-field strain on the surface of the gage length of the test

specimen throughout the test. The tests were conducted under crosshead

displacement control at a rate of �̇� = 0.0127 𝑚𝑚/𝑠 and the load (F) was

recorded every 4 seconds using a 14bit Win600 Digital Oscilloscope.

Assuming plastic incompressibility, the true strain and true stress are

0

0 0

lnl

l

F l

A l

(3-2)

where l is the current length, l0 is the initial length, and A0 is the initial cross-

sectional area.,

3.2.1 Characterization of ES-1 under Uniaxial Compression

At the AWEF, compression tests were conducted in accordance with

ASTM E9 using standard cylindrical specimens of dimensions shown in

Figure 3-3 with lubricant on each face to minimize radial forces at an average

strain rate of 0.000333 s-1. Each test was repeated four times.

Page 67: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

67

Figure 3-3. Quasi -static cylindrical compression test specimens.

Note the consistency in the stress-strain response displayed in Figure 3-4

for all three ES-1 materials.

7.62 mm

7.62 mm

Page 68: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

68

Figure 3-4. Quasi-static round bar compression test stress-strain results A) forged specimens, B) cast and HIP’d specimens, and C) cast

specimens.

A

B

C

Page 69: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

69

As expected based on the microstructural analysis, the stress-strain

behavior of the cast and cast and HIP’d materials in uniaxial compression are

similar, and the forged material is the strongest (where the stress-stain curves for

the three ES-1 materials are superposed in Figure 3-5). For example, at 13%

strain, the average compressive stress for the forged material is 2063 MPa,

compared to 1937 MPa and 1961 MPa for the cast and cast and HIPd materials,

respectively.

Figure 3-5. Comparison of the quasi-static compressive stress-strain response for the ES-1 materials.

Page 70: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

70

For each ES-1 material, in accordance with ASTM E9, the yield point, the

beginning of nonlinear stress-strain behavior was determined consistently using

0.2% strain offset. The yield stress σy of each material are given in Table 3-2.

Table 3-2. Yield stress and quasi-static compression test data. Manufacturing Process Specimen σy (MPa)

Forged X29978 1616 X29979 1366 X29980 1503 X29981 1496 Cast and HIP’d X29876 1448 X29877 1340 X29878 1189 X29879 1336 Cast X29928 1395 X29929 1403 X29930 1345 X29931 1340

From the stress-strain curves, both Swift (see Eq. (1-17)) and Voce

hardening laws (see Eq. (1-18)) were identified (see also Figure 3-6). The

numerical values of the parameters involved in each law are given in Table 3-3.

Table 3-3. Coefficients involved in Swift and Voce Hardening Laws for ES-1. Manufacturing Process k

(MPa) n

Av (MPa)

Bv

(MPa) Cv

Forged 2436 0.0708 2044 504.5 63.33 Cast and HIPd 2465 0.0921 1934 556.4 66.20 Cast 2349 0.0777 1934 548.5 67.48

Note that hardening under compression of each ES-1 material is best

characterized using Swift law.

Page 71: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

71

Figure 3-6. Hardening of ES-1 under compression according to Swift and Voce A) forged ES-1, B) cast and HIP’d, and C) cast ES-1.

Page 72: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

72

3.2.2 Quasi-static Characterization of ES-1 in Uniaxial Tension

At the AWEF, quasi-static, uniaxial tension tests were conducted in

accordance with ASTM E8 standard using standard dog bone specimens

illustrated in Figure 3-7 at an average strain rate of 0.000365 s-1. For each

ES-1 material, the yield stress σy,, the stress at which necking is observed σu,

and the strain-to-necking (STN) are reported in Table 3-4.

Figure 3-7. Schematic of the quasi-static round tensile test specimen.

Table 3-4. Summary of quasi-static tension test data. Manufacturing Process Specimen σy

(MPa) σu

(MPa)

STN

Forged X29938 1601 1835 0.050 X29939 1640 1829 0.052 X29940 1456 1841 0.050 X29941 1465 1838 0.054 Y08613 1535 1782 0.041 Y08614 1506 1758 0.043 Cast and HIP’d X29836 1483 1733 0.047 X29837 1455 1710 0.044 X29838 1411 1736 0.049 X29839 1431 1726 0.049 ES1_C_H1 1425 1646 0.036 ES1_C_H2 1481 1681 0.052 Cast X29888 1428 1711 0.062 X29889 1456 1716 0.074 X29890 1447 1741 0.093 X29891 1450 1722 0.035 ES1_C_1 1400 1644 0.030 ES1_C_2 1390 1652 0.021 ES1_C_3 1407 1649 0.027

As for uniaxial compression loadings, the forged ES-1 exhibits higher

yield stress than the other ES-1 materials, the increase factor being of 1.1, the

uniaxial tensile response of the cast and cast and HIP ES-1 materials being

similar (see also Figure 3-8). The Swift and Voce hardening laws were also

6.35 mm

25.4 mm

31.75 mm

6.35 mm

Page 73: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

73

identified based on the uniaxial tensile results. The numerical values of the

respective coefficients are reported in Table 3-5 and capture fairly well the

stress-strain curve until the onset of necking for all three manufacturing

procedures (Figure 3-8). Furthermore, the initial tensile yield stress of the forged,

cast, and cast and HIP’d ES-1 under uniaxial tensile loading corresponds to the

stress documented by Boyce and Dilmore (2009), Dilmore and Ruhlman (2009),

Van Aken et al. (2014), and Lynch (2011) at similar quasi-static strain rates. The

STF of the materials produced by the three manufacturing processes: forged,

cast and HIP’d, and cast were 5.15 %, 4.75 %, and 7.77%, respectively, similar

to that reported by other authors including van Aken et al. and Lynch. However,

the STF reported by Boyce and Dilmore was much higher than that of any of the

manufacturing processes examined in this study possibly due to additional heat

treatment processing.

Table 3-5. Hardening law parameters tensile round specimens. Manufacturing Process k

(MPa) n

Av (MPa)

Bv

(MPa) Cv

Forged 2272 0.0621 1836 307.1 110.5 Cast and HIPd 2171 0.0655 1722 303.5 139.1 Cast 2106 0.0578 1717 268.0 137.3

Page 74: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

74

Figure 3-8. Stress strain curves for ES-1 quasi-static tension test results of forged, cast and HIP’d, and cast no HIP specimens.

The fracture surfaces of the round tensile test specimens are generally the

typical cup and cone displayed in Figure 3-9 similar to the ES-1 fracture surfaces

observed by Boyce and Dilmore (2009). The cross section of the fractured

specimens remained circular, further evidence of the isotropy of ES-1

irrespective of manufacturing process.

Page 75: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

75

Figure 3-9. Fracture surface of the round specimens following tensile tests A) forged, B) cast and HIPd, and C) cast.

Several recent studies have demonstrated that ductility measurements are

inversely proportional to maximum cluster pore volume and pore fraction at

fracture (Susan et al., 2015, Foley et al., 2016). Consistent with these findings,

note the large porous regions in the cast ES-1 fracture surface and the lack

thereof in the cast and HIP’d ES-1 material.

The purpose of completing quasi-static tension and compression tests

was to establish the symmetry properties of ES-1. Like other BCT martensitic

steels discussed in Section 1.2, Eglin steel demonstrates higher flow stress in

uniaxial compression as compared to tension. Furthermore, the tension-

compression asymmetry evolves with the equivalent plastic strain. At initial

yielding, the tension-compression asymmetry is not pronounced but, the

asymmetry becomes significant at the ultimate stress. It is important to note that

the tension-compression asymmetry is reported only in the range of equivalent

plastic strain where the deformation in the flat tensile specimen is homogeneous

(i.e. before necking). The tension-compression asymmetry ratio, t/c for several

values of equivalent plastic strain is reported in Table 3-6. The cast specimens

Page 76: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

76

did not reach a plastic strain of 4% therefore the tension-compression asymmetry

ratio was not calculated or reported.

Table 3-6. Tension-compression asymmetry ratio of ES-1 by plastic strain. Manufacturing Process/Plastic Strain 0.2% 2% 4%

Forged 1.0169 0.9249 0.8924 Cast and HIPd 1.0936 0.9375 0.8859 Cast 1.0205 0.9015 N/A

The asymmetry ratio of ES-1 at 4% strain is similar to the strength ratio of

SAE 4340 0.90T

C

and other martensitic steels reported by Leslie and Sober

(1967).

3.2.3 Influence of Specimen Geometry on the Mechanical Response in Uniaxial Tension

In order to study the influence of the geometry of the specimen on the

mechanical response in uniaxial tension, tension tests were also carried out

using quarter scale standard flat pin-loaded specimens shown in Figure 3-10.

Figure 3-10. Schematic of the quasi-static (flat) pin-loaded tensile test specimen.

Figure 3-11 allows a comparison between the stress-strain response of

cylindrical specimens and flat pin-loaded specimens for each type of processing.

It is worth noting that, as predicted, the stress-strain response prior to strain

localization was similar for both specimen geometries: round and flat. Though,

the STF of the flat specimens was about half that of the round specimens. It is

hypothesized that the decrease in STF is at least partially a result of the subscale

50.8 mm

12.7 mm

3.175 mm

0.156 mm

3.175 mm

14.29 mm

12.7 mm 12.7 mm

Page 77: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

77

specimens containing relatively larger flaws. The hardening law parameters are

presented in Table 3-7.

Table 3-7. Hardening law parameters flat tension specimens. Manufacturing Process k

(MPa) n

Av (MPa)

Bv

(MPa) Cv

Forged 2157 0.06 1774 260.2 113.6 Cast and HIPd 1983 0.0491 1664 205.8 125.9 Cast 2210 0.0707 1653 423.8 290.1

Page 78: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

78

Figure 3-11. Quasi-static tension characterization stress-strain response A) forged, B) cast and HIP’d, and C) cast no HIP specimens.

Page 79: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

79

Although the linear variable differential transformer (LVDT) can provide

strain measurements, DIC measurement was used to monitor the evolution of the

local strain. To gain understanding of the plastic deformation of the ES-1, two

cameras were disposed to monitor, independently, the evolution of the strain in

the gage length. The first camera shown in Figure 3-12 associated with 2-

dimensional DIC measurements captured the evolution of strain occurring in the

face of the specimen, i.e. the width strain (width) and axial strain (axial), while the

second camera monitored the evolution of strain occurring in the side of the

specimen, i.e. the thickness strain (thickness) and axial strain (axial). Both cameras

were synchronized using a delay generator to capture images at the same time

and hence the same axial displacement.

Figure 3-12. Camera layout to capture 2-D measurements of the strain evolution on the face and the side of the flat specimens during quasi-static tensile testing.

Lankford coefficients, also known as r-values, are a measure of the

thinning resistance of a sheet or plate during forming and are given as the ratio of

width to thickness strain rate width

thickness

r

.

Taking advantage of the experimental setup that allows the independent

monitoring of the evolution of the width strain and the thickness strain presented

Page 80: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

80

in Figure 3-13, there is no need to adopt the usual hypothesis of plastic

incompressibility. Using a linear regression, the Lankford coefficient of one was

obtained using direct measurements of the width and thickness strain rates

meaning that the material is indeed isotropic.

With an understanding of the quasi-static behavior of ES-1, dynamic

SHPB compression tests were conducted to determine the rate-sensitivity as

discussed in the ensuing section.

Page 81: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

81

Figure 3-13. Width and thickness strain of flat specimens under uniaxial tension A) forged, B) cast and HIP’d, and C) cast no HIP specimens.

Page 82: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

82

3.3 Dynamic Experimental Characterization of Eglin Steel

The Split Hopkinson Pressure Bar (SHPB) or Kolsky bar is a test

apparatus for material characterization at high strain rates on the order of 103 s-1

since it was developed by Kolsky (1949). A diagram of a Kolsky bar included in

Figure 3-14 shows the striker bar, incident bar, and transmitted bar aligned in

series.

Figure 3-14. Schematic SHPB with an illustration of the propagating strain waves.

In the operation of a SHPB, the striker bar is put into motion to impact the

incident bar creating a compressive strain wave composed of a spectrum of

frequencies which backward propagates the length of the striker bar and forward

propagates the incident bar. Within the striker bar, the compressive wave is

reflected at the free end creating a tensile wave forward propagating back though

the length of the striker bar (Ls). Upon the arrival of the tensile wave at the

impact face of the striker bar, the striker bar is pulled away from the incident bar

concluding the compressive wave in the incident bar with a final wavelength of

2Ls/C0 where C0 is the longitudinal wave speed in the striker bar. When the

compressive wave impinges upon the specimen, the strain wave is both reflected

and transmitted at both interfaces creating a tensile reflected pulse that backward

propagates the incident bar and a diminished compressive wave that continues

Page 83: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

83

forward propagation through the transmitted bar. Strain gages are adhered to

the surface of both the incident and transmitted bars to collect engineering strain

at a distance xi and xt, respectively, from the specimen throughout the dynamic

compression test. The placement of the strain gages is critical to eliminate

edged effects and avoid undesirable interaction of the reflected waves at the

strain gage locations. One dimensional wave theory is generally used in the

analysis of a SHPB to determine the average, uniaxial stress and strain within

the cylindrical test specimen throughout the test. Using the free body diagram of

a section of length dx of the incident or transmitted bar under stress shown in

Figure 3-15, the equation of motion becomes:

2

2

uA A Adx

x t

(3-3)

Where

ρ is the density of the incident and transmitted bars,

u is the longitudinal displacement of particles within the bars, and

A is the cross sectional area of the incident and transmitted bars.

Figure 3-15. A free body diagram of a portion of the SHPB of length dx and cross sectional area A.

Simplifying and assuming (1) both the incident and transmitted bars are

not plastically deformed (i.e. Hooke’s law is applicable) and (2) the bars are long

enough to assume uniaxial strain, the one-dimensional wave equation introduced

by Jean le Rond d’Alembert is:

Page 84: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

84

2 2

2 2

u E u

t x

(3-4)

where E is the Young’s modulus of the bars.

By introducing the longitudinal wave speed, defined as 0

EC

the

general solution to the differential equation is:

0 0u f x C t g x C t (3-5)

for displacement forward (f) and backward (g) propagating waves. For the

forward propagating incident and transmitted waves and the backward

propagating reflected wave, the displacement at the interfaces of the specimen

(u1 and u2) can be calculated:

1 0

0

2 0

0

t

i r

t

t

u C dt

u C dt

(3-6)

Where

εi is the strain measured as the incident wave initially passes the strain gage on the incident bar,

εt is the strain measured in the transmitted bar, and

εr is the strain measured on the incident bar after the wave has reflected at the specimen interface.

Notice the data are time-shifted, such that the strain can be integrated

from time, t=0. Remember that the strain gages are positioned xi and xt away

from the specimen interface creating a voltage time graph like the one shown in

Figure 3-16. The physical interpretation is that the incident strain is measured

Page 85: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

85

first when the incident wave reaches the first strain gage before it ever reaches

the first specimen interface. It is assumed that the incident strain wave does not

change before it reaches the first specimen interface. Therefore, it is forward

propagated to the first interface with the specimen simply by adding the time it

takes for the wave to propagate the distance from the incident strain gage to the

first specimen interface, incident time 0

ii

xt

C

. After the incident stain is

measured and the wave propagates to the first specimen interface, the amplitude

of the incident wave is then transmitted and reflected at both interfaces of the

specimen. The portion that is transmitted through the specimen, propagates

down the transmitted bar until it reaches the transmitted strain gage. Therefore,

the transmitted time 0

tt

xt

C

is subtracted from the time associated with the

transmitted wave. Finally, the portion of the wave that was reflected at the

interfaces with the specimen is then measured when it returns to the incident

strain gage. Therefore, incident time, is subtracted from the time associated with

the reflected wave measurement. Finally, the new time, which should coincide

with the time that each wave was interacting with the specimen interfaces is reset

at 𝑡 = 0. This model ignores any changes in the wave as it propagates between

the incident strain gage and the first specimen interface and between the second

interface and the transmitted strain gage and neglects the time the wave travels

within the relatively short test specimen and is only an accurate assumption if the

SHPB has been designed to minimize dispersion, distortion of the wave as it

propagates the SHPB system due to phase speeds of the frequency components

Page 86: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

86

of the composite strain wave, to a negligible amount. Distortion is minimized

when the wave length is much larger than the SHPB diameter as explained by

Davies (1948).

Page 87: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

87

Figure 3-16. Raw data from the dynamic SHPB characterization A) forged, B) cast and HIP’d, and C) cast specimens.

Page 88: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

88

The average strain (εs) in the specimen of length (L) is calculated based

on the displacement of the two interfaces.

02 1

0

t

s i r t

Cu udt

L L

(3-7)

Assuming a planar impact at both specimen interfaces, the stress at each

interface can be calculated using Hooke’s Law and the total strain at each

interface 1 2;i r t . The specimen free body diagram is shown in Figure

3-17.

1

2

i r

s

t

s

E A

A

E A

A

(3-8)

Where As is the cross sectional area of the specimen.

The average stress within the specimen can be estimated based on the

stress at each of the specimen interfaces.

1 2

2 2

i r t

s

s

E A

A

(3-9)

Figure 3-17. The SHPB test specimen subjected to forces at the incident and transmitted bar interfaces.

The materials lab at the University of Florida (UF) Research and

Engineering Education Facility (REEF) contains a SHPB made of three 0.75-in

diameter Vascomax-300 steel bars aligned in series: striker bar, incident bar, and

Page 89: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

89

transmitted bar, on a rigid optical table shown in Figure 3-19. Vascomax-300 has

a Young’s modulus of 190 GPa and a density of 8,000 kg/m3.

Figure 3-18. UF REEF test equipment for dynamic SHPB characterizations A) HDO 8058 500 MHz high-definition oscilloscope, B) amplifier, and

C) SHPB.

3.3.1 Dynamic Stress Strain Response of ES-1 under Uniaxial Compression

The incident, transmitted, and reflected strain measurements collected at

500 MHz during the SHPB tests of forged, cast and HIP’d, and cast specimens

with a length and diameter of 5.08 mm are included in Figure 3-19 A, B, and C,

respectively, to provide insight into the duration and dispersion of the strain

waves.

Page 90: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

90

Figure 3-19. Strain waves at the incident and transmitted strain gages A) forged, B) cast and HIP’d, and (C) cast specimen.

Page 91: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

91

The stress-strain response of ES-1 under dynamic, compressive loading is

shown in Figure 3-20. The forged material demonstrates slightly higher flow

stress than the cast and HIP’d and cast no HIP materials and thus lower plastic

strain as the same kinetic energy (i.e. mass and impact velocity of the striker bar)

was introduced into the system for each test.

Figure 3-20. Stress strain response of forged, cast and HIP’d, and cast no HIP ES-1 during SHPB tests.

The deformed length of the post-test specimens was used to validate the

one-dimensional analysis. A photograph of the forged ES-1 cylindrical test

specimen initially and with 16% plastic strain following a SHPB dynamic

compression test included in Figure 3-21 proves the one-dimensional analysis

was accurate in the SHPB at the UF REEF. Additionally, to correct the wave

form for dispersion (Gong et al., 1990), a Fourier transform was used to

Page 92: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

92

transform the data into the frequency domain, a variation of phase velocity using

the Pochhammer-Chree (Pochhammer, 1876 and Chree, 1889) solution was

introduced, and finally a reverse Fourier transform back to the time domain was

completed to confirm the UF REEF SHPB system permits very little dispersion.

Figure 3-21. Forged Eglin steel cylindrical SHPB test specimens A) before and B) after dynamic compression test.

Figure 3-22, the stress-strain response of the forged, cast and HIP’d, and

cast no HIP materials, permits comparison of the quasi-static and dynamic

behavior of Eglin steel.

Page 93: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

93

Figure 3-22. Comparison of quasi-static and dynamic stress-strain response A) forged, B) cast and HIP’d, and C) cast no HIP ES-1.

Page 94: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

94

Because of the difficulty in determining flow stress for small percentages

of plastic strain (Sharpe and Hoge, 1972, and Yadav et al., 1995) and to

eliminate the inertial effects in the dynamic stress-strain curves, the quasi-static

and dynamic flow stresses were compared for strains greater than 4% to

determine the forged dynamic flow stress is on average 1.10 times the quasi-

static stress for the same values of strain; the cast and HIP’d is 1.14; and the

cast no HIP is 1.12. These dynamic increase factors are nearly equal to the

dynamic increase factor of 1.18 reported by Johnson and Cook (1983) following

testing of SAE 4340. So, the dynamic increase factor can be used to solve the

dynamic portion of the isotropic Johnson Cook presented previously in Section

1.2.

Where the material constants, presented in Table 3-8, are similar to those

presented by Dilmore (2009) after characterizing the behavior of ES-1 at various

strain rates using a servohydraulic system with custom built load cells.

Table 3-8. Johnson and Cook material constants for forged, cast, and cast and HIP’d ES-1.

Manufacturing Process A (MPa) B (MPa) n C

Forged 1430 1142 0.2525 0.0124 Cast and HIP’d 1414 1184 0.2344 0.0065 Cast 1445 1149 0.2406 0.0069

Referring back to the literature, the initial yield stress of AF-1410, AISI

4340 and ES-1 observed by Last (1996) is presented in Figure 3-23 along with

the cast ES-1 experimental results. The initial yield stress of ES-1 reported by

Last corresponds well with the cast ES-1.

Page 95: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

95

Figure 3-23. Initial yield stress-strain-rate dependence of AF-1410, AISI 4340, and ES-1 observed by Last in comparison to the cast material.

As shown in Figure 3-24, during the SHPB test, the strain rate increased

rapidly to a maximum of over 1,500 s-1 and declined to about 500 s-1 with an

average strain rate of approximately 700 s-1. To minimize the inertial effects,

forged ES-1 specimens were tested in a SHPB with a pulse shaper adhered to

the leading end of the incident bar as discussed in the following Section, 3.3.2.

Page 96: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

96

Figure 3-24. Strain rate in forged, cast and HIP’d, and cast no HIP ES-1 during SHPB tests.

3.3.2 Dynamic Characterization of ES-1 at Constant Strain Rates

Without a pulse shaper, the strain rate in the test specimens increased

rapidly when the incident wave impinged the specimen and decayed

exponentially throughout the duration of the incident wave. Therefore, pulse

shapers, pieces of material placed on the leading end of the incident bar using

lubricant, were used to absorb the initial impact energy and induce an

approximately constant strain rate in the test specimen. The design of the pulse

shaper is critical to obtaining nearly constant strain rates and dynamic

equilibrium. Frew et al. (2005) present analytical models and experimental data

that demonstrate a broad range of incident pulses and can be obtained by

varying the design of the pulse shaper. For the Eglin steel specimens, two

different pulse shapers were designed to test at strain rates of approximately 500

s-1 and 2,000 s-1. The 500 s-1 pulse shaper was made from an 1/8-in thick,

Page 97: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

97

annular piece of 4130 steel with four 0.012-in thick circular pieces of annealed

copper equally spaced around the center as shown in Figure 3-25A. The 2,000

s-1 pulse shaper was made from a 1/8-in thick, annular ring of 4130 steel, a

0.012-in thick annular ring of annealed copper and a circle of 0.01-in thick Teflon

as shown in Figure 3-25B.

Figure 3-25. Dynamic pulse shapers made of copper, steel, and Teflon A) 500 /s and (B) 2000 /s.

In order to obtain similar plastic deformation despite the energy absorbed

in the pulse shaper, the SHPB test was designed with an increased pressure of

55 psi and a 2-ft striker bar. Three SHPB tests were conducted on the forged

specimens using the pulse shapers: two at 500 s-1 and one at 2000 s-1. Notice

Page 98: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

98

the more constant strain-rate behavior of the 500 s-1 pulse shaper presented in

Figure 3-26.

Figure 3-26. Strain rate of forged test specimens without a pulse shaper, with the first pulse shaper, and with the second pulse shaper.

The stress-strain response of forged ES-1 under constant strain rates of

500 s-1 and 2000 s-1 obtained using the pulse shapers are presented alongside

the forged ES-1 tested in the SHPB without a pulse shaper in Figure 3-27. The

flow stress of the forged material tested at the highest strain rate 2000 s-1 is

proportionately greater than at 500 s-1 and 700 s-1.

Page 99: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

99

Figure 3-27. Stress strain curves of forged specimens without a pulse shaper, with the first pulse shaper design and the second pulse shaper design.

3.4 Cyclic Experimental Characterization of ES-1

Additional 25.4-mm long, round specimens were tested under cyclic

loading with an MTS 25-ton capacity test frame and load cell in the materials lab

at the UF REEF. The load frame, high-speed camera, and associated equipment

are shown in Figure 3-28.

Page 100: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

100

Figure 3-28. UF REEF materials lab for cyclic characterization of ES-1 A) test frame and load cell, B) high-speed camera, and C) installed

specimen.

Prior to testing, the specimens were cleaned with acetone and marked

with a vertical line for alignment of the high-speed camera and targets evenly

spaced at 1 mm to track local displacement with the high-speed cameras. The

specimens were subjected to uniaxial tension followed by uniaxial compression

under completely reversed crosshead displacement control at a frequency of 0.1

Hz. The multi-test approach described in STP-465 was initialized with 0.2%

strain amplitude for 50 cycles to reach stability. Next, the strain amplitude was

Page 101: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

101

then increased by 0.2% for each subsequent iteration. Images of the gage

length of the test specimen illuminated by a stationary LED array were recorded

using an Allied high-speed camera and 25 mm lens. The nested hysteresis

loops at various values of maximum plastic strain of the forged, cast, and cast

and HIP’d ES-1 materials are included in Figure 3-29.

Figure 3-29. Hysteresis loops of ES-1 under completely reversed displacement A) forged, B) cast and HIP’d, and C) cast.

The absolute uniaxial flow stresses in tension and compression at the

extremes of the elastic range under tensile plastic strain and compressive plastic

strain of forged ES-1 calculated using the maximum values or 0.1% offset as

required are presented in Figure 3-30. A large variability in flow stress was

observed depending on the method used to determine the flow stress (i.e.

proportionality limit or percent offsets between 0.01 and 0.2%). Nevertheless,

note the Bauschinger effect, softening in compression under tensile strain and

vice versa.

Page 102: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

102

Figure 3-30. Uniaxial absolute flow stress of ES-1 at the limits of the elastic region under cyclic loading by plastic strain.

The loci of the maximum values of stress and strain from all the stable

hysteresis loops together form the kinematic stress-strain response of forged ES-

1. The kinematic stress-strain response is shown in Figure 3-31 alongside the

quasi-static response for comparison. Referring back to the literature, many

materials including martensitic steels undergo kinematic softening under cyclic

loading as demonstrated by a comparison of the quasi-static and kinematic

stress-strain curves presented by Landgraf (1970) included in Chapter 1 of this

dissertation. However, the kinematic softening of forged, cast and cast and

HIP’d ES-1 is small for the range of true strain, contrasting the extreme softening

of SAE 4340 at initial yielding.

Page 103: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

103

Figure 3-31. Monotonic and kinematic stress-strain curves of Eglin steel A) forged, B) cast and HIP’d, and C) cast no HIP.

Page 104: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

104

3.5 Summary of the Experimental Characterization of ES-1

The experimental characterization of Eglin steel revealed the forged

material flow stress was higher than the cast and cast and HIPd materials by a

factor of 1.1 irrespective of the sign of the loading and strain rate. The flow

stress of each material under quasi-static conditions was accurately described

using Swift hardening law. Additionally, the material regardless of manufacturing

route exhibited isotropy with comparable width and thickness strain rates within

the flat tension test specimens (r=1). The SHPB tests divulged a slight strain-

rate dependence well characterized by the Johnson Cook (1983) constitutive

model. Two characteristics were discovered during the cyclic testing: (1) Eglin

steel softens under repeated load reversals, and (2) the material exhibits a

pronounced Bauschinger effect. These material characteristics and associated

material parameters were used to develop the theoretical elastic-plastic model

presented in Chapter 4 of this dissertation.

Page 105: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

105

CHAPTER 4 ELASTIC-PLASTIC MODEL FOR EGLIN STEEL

The main features revealed by the experimental data are used to develop

an appropriate theoretical model valid for general 3D loadings. As UHSS and

ES-1 steels have a well-defined yield point as seen in the stress-strain response

under uniaxial loading (see Chapter 3). To describe both ineslastic behavior and

the rate effects revealed by the SHPB tests and data, viscoplasticity theory was

used to model ES-1. In such theory, an appropriate yield criterion and hardening

law need to be defined.

, , 0p p pF Y σ σ (4-1)

As already mentioned, in the six dimensional stress space, the yield

criterion defines a surface such that for any state interior to it, the material

behavior is elastic. After initial yielding, the self-similar expansion of the

subsequent yield surfaces is usually considered to be governed by the equivalent

plastic strain. In this dissertation, isotropic hardening includes not just expansion

of the yield surface, but also distortion of the yield surface that preserves the

isotropic nature of the deformation. Rate effects can be introduced by either

considering an overstress approach or by considering that hardening is rate-

dependent. This latter approach is used.

4.1 Development of the Yield Criterion for ES-1

Several yield criteria and associated effective stresses have been

proposed in the literature (see Section 1.4), of which, most used are von Mises

(1913) and Drucker (1949), respectively.

Page 106: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

106

2

13 2 62 3

3J

J cJ

(4-2)

where 𝐽2 and 𝐽3 are the second and third invariants of the deviatoric stress, s.

The deviatoric stress is the stress tensor less the hydrostatic stress tensor. The

second and third invariants of the deviatoric stress are the coefficients of the first

and zeroth order terms of the characteristic equation of the deviatoric stress

eigenvalue problem expressible in terms of the principal stress (1 2 3, , ),

principal deviatoric stress (1 2 3, ,s s s ), the Cauchy stress components

11 22 33 12 23 13, , , , , , or the deviatoric stress tensor components

11 22 33 12 23 13, , , , ,s s s s s s . The invariants are so named because their expression is

the same irrespective of the coordinate system. Let us recall that, in terms of

principal stresses 1 2 3, , :

2 2 2

2 1 2 2 3 3 1

1

6J

. (4-3)

While for general loadings, the expression is:

2 2 2 2 2 2

2 11 22 33 11 22 22 33 11 33 12 23 13

1 1

3 3J (4-4)

Likewise, the third invariant of the deviatoric stress is:

3 3 3 2 2 2

3 1 2 3 1 2 3 2 1 3 3 1 2 1 2 3

2 1 4

27 9 9J (4-5)

3 3 3 2 2 2

3 11 22 33 11 22 33 22 11 33 33 11 22 11 22 33

2 22

23 131211 22 33 11 22 33 11 22 33 12 23 13

2 1 4

27 9 9

2 2 2 23 3 3

J

(4-6)

Page 107: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

107

Because von Mises (1913) and Drucker (1949) are functions of the

deviatoric stress invariants, martensitic steel material properties of pressure

insensitivity and isotropy are easily captured with these yield criteria. However,

these yield criteria are even functions of stress and therefore cannot capture the

experimentally observed tension-compression asymmetry. In the experimental

characterization, ES-1 exhibits higher yield stress in compression than tension

for a wide range of plastic strain. Therefore, a yield criterion that accounts for the

tension-compression asymmetry such as Cazacu and Barlat (2004) yield criterion

was used to describe the elastic-plastic behavior of the ES-1 steel alloys.

Similarly to Drucker (1949), this yield criterion is expressed as a function of the

invariants of the stress deviator, but contrary to the Drucker (1949) (see Eq.

(4-2)) which is an even function, the Cazacu and Barlat (2004) model is an odd

function of stresses. The effective stress associated with this yield function is:

13 322 3 YJ cJ

(4-7)

where c is an asymmetry parameter and Y is the shear yield stress. In Cazacu

and Barlat (2004) it was shown that c can be expressed using the flow stresses

in tension (T) and compression (C) as:

3 3

3 3

3 3

2

T C

T C

c

(4-8)

The main feature of the Cazacu and Barlat (2004) criterion is that it is an

odd function, which allows describing the tension-compression asymmetry using

the third invariant of the deviatoric stress. In the principal stress space, the

Page 108: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

108

Cazacu and Barlat (2004) yield surface is an infinite cylinder with generator, the

hydrostatic axis. In the deviatoric plane, normal to the hydrostatic axis, the cross

section of the yield surface is a rounded “triangle” for a nonzero asymmetry

parameter 0c or a circle for 0c .

In this research a different method for determining this parameter and the

evolution of yield surfaces is proposed. Before discussing these aspects, the flow

rule and application of the criterion to description of the torsional response is

presented.

4.1.1 Associated Flow Rule

Drucker (1951) used the non-negative increment of plastic work to prove

normality (see Figure 4-1A).

0Pdσ :dε (4-9)

At the limiting case when the increment of stress is tangent to the yield

surface, the increment of plastic strain must be normal to the yield surface (i.e.

the increment of plastic work is zero).

The material being isotropic, it is sufficient to describe the plastic flow in

the coordinate system associated with the principal values of stress. The flow

rule for Cazacu and Barlat (2004) is:

32

2 3

P

i

i i

JJF F

J J

(4-10)

1

3222

32

232 3 322 3

3

2

3

i i

i i

JJJ c

JJF F

J J

J cJ

(4-11)

Page 109: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

109

Let us recall that in 1951 Drucker proved that the work from a cycle of

loading and unloading between an arbitrary elastic stress and stress resulting in

an increment of plastic strain is positive only if the entire yield surface is on one

side of the hypertangent plane at the point of plastic stress.

0 Pσ σ* dε . (4-12)

Figure 4-1. Illustration of properties of yield surfaces with an associated flow rule A) normality and B) convexity.

The condition of positive work is satisfied at every point for a convex

instantaneous yield surface (see Figure 4-1B). For the entire yield surface to

remain convex the Hessian must be positive semi-definite for any fixed value of

hardening. The Hessian is positive semi-definite if the three eigenvalues at all

points of the yield surface are all non-negative.

222 2

32

2 2

2 3

0i j i j

JJF FH

J J

(4-13)

The local maximums of the convex yield surface in the deviatoric plane

occur along both the tensile and compressive meridians (projections of the

positive and negative principal stress half-axes, respectively, in the deviatoric

Page 110: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

110

plane) where the determinant of the Hessian, and therefore an eigenvalue, are

both zero. Consequently, the Hessian is positive semi-definite if both the sum

and the product of the remaining, conjugate eigenvalues are positive (i.e. the first

and second invariants of the Hessian are both positive).

For the Cazacu and Barlat (2004) yield criterion, it is important that

convexity of the yield function be reinforced to find the acceptable range of the

asymmetry parameter.

1 1

32 2 22 22 2

23 322 3

1 1

3 32 22 22 2

53 322 3

3 3

4 2

3

3 32

2 2

9

i j i j i j

i i j j

JJ J JJ J c

H

J cJ

J JJ JJ c J c

J cJ

. (4-14)

It can be easily seen that:

2ij

ij

Js

(4-15)

2

2 1

3ik kl ij kl

ij kl

J

, (4-16)

and that:

32

2

3ik kj ij

ij

Js s J

(4-17)

2

3 2

3ik jl jl ik ij kl kl ij

ij kl

Js s s s

. (4-18)

Page 111: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

111

In Cazacu and Barlat (2004), it was shown that the convexity requirement

results in an acceptable range of the asymmetry parameter 3 3 3 3

,4 4

c

.

4.1.2 Unusual Behavior in Torsion Revealed by Cazacu and Barlat (2004) Criterion

The Swift effect, put into evidence by Swift (1947), is the axial deformation

of a thin-walled cylinder during monotonic free-end torsion. This phenomenon

was previously attributed to anisotropy and/or its evolution. In 2013, Cazacu et

al. (2013) gave a different explanation. It was put into evidence that there is a

correlation between a material asymmetry and Swift effects. A theoretical

analysis was done using the CBP-06 yield criterion which accounts for tension-

compression asymmetry using the parameters, k and a.

1 1 2 2 3 3

a a a

s ks s ks s ks F (4-19)

Revil et al. (2014) conducted a finite element analysis using an elastic-

plastic model based on the anisotropic form of CBP-06 and described the

behavior of thin-walled cylinders of OFHC copper, high-purity α-titanium, and

AZ31-Mg alloy under torsion. It was confirmed that axial elongation occurs in

materials with a tension to compression flow stress ratio greater than one, and

contraction of materials with a ratio less than one.

In this work, analysis of the same problem is done using the Cazacu and

Barlat (2004) yield criterion. Let the applied loading be given by:

0 0

0 0

0 0 0

σ . (4-20)

Page 112: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

112

Hence its deviator is:

0 0

0 0

0 0 0

s . (4-21)

Under this state of stress, the second invariant of the deviatoric stress is

equal to the shear stress and the third invariant is zero.

2 2 2 2

2 1 2 3

3 1 2 3

1

2

0

J s s s

J s s s

(4-22)

Substituting these values in the associated flow rule, we obtain:

1

22

2 11 3 3 33 2

2 3

30 0

2 33

32 0 02 3

3 3 223 3 2

0 03

c

J cd c

d d

ccc J cJ

Pdε

32 ∂J∂J∂F ∂σ ∂σ

∂σ (4-23)

Note that according to Eq. (4-23), if the asymmetry parameter, c, is zero,

then pure shear deformation occurs (i.e. no axial deformation). If the asymmetry

parameter is negative the cylinder will elongate, where if c is greater than zero, it

will contract. A numerical solution using finite element analysis will also be

presented in the next chapter.

4.2 Asymmetric Hardening

As already mentioned, based on the classical definition, isotropic

hardening is modeled, typically, by a function of equivalent plastic strain

parameterized using the quasi-static uniaxial tensile characterization. The

subsequent yield surface, symmetric about the origin, therefore enlarges to pass

through the subsequent, uniaxial tensile flow stress. Therefore, a similar

Page 113: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

113

increment of hardening is added to each point on the yield surface. The initial

yield surface and two subsequent yield surfaces using classical isotropic

hardening for forged ES-1 are presented in Figure 4-2.

Figure 4-2. Cazacu and Barlat initial and two subsequent yield surfaces using classical isotropic hardening of forged ES-1 in the deviatoric plane.

In this research a new description of isotropic hardening is proposed.

Here, it is hypothesized that isotropic hardening is a result of both a change in

size and shape of subsequent yield surfaces that does not destroy the isotropy.

In this research, the deformation of the subsequent yield surface is described

using the asymmetry parameter as a function of the second and third invariants

of plastic strain; and the enlargement is described using the hardening under

pure shear strain. With the evolution of asymmetry of the yield surface

dependent on the second and third invariants of plastic strain, the subsequent

Page 114: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

114

yield surfaces vary based on the strain path. For example, the subsequent yield

surface following uniaxial tension may be different from the subsequent yield

surface following uniaxial compression.

For quasi-static monotonic loading, hardening is generally considered to

be a function of plastic strain though some models do include other variables like

fading strain history and other internal variables. However, for a plastically

incompressible, isotropic material, the hardening can be described using both the

second and third invariants of the total plastic strain, just as the yield surface in

Haigh-Westergaard space is defined by the second and third invariants of the

deviatoric stress (i.e. the hardening is a function of the second order plastic strain

tensor that can be completely described using both non-zero invariants). The

strain invariants, similar to the deviatoric stress invariants, are the coefficients of

the first and zeroth order terms of the eigenvalue problem in strain expressible as

a function of the components of the strain tensor 11 22 33 12 23 13, , , , , or the

principal strains 1 2 3, , . As a result of plastic incompressibility, the plastic

strain is deviatoric, that is the first invariant of plastic strain is identically zero (i.e.

1 1 2 3 0J ).

2 2 2 2 2 2 2 2 2

2 1 2 3 11 22 33 12 23 13

1 1

2 2J (4-24)

3 1 2 3

3 3 3 2 2 2

11 22 33 12 11 22 23 22 33 13 11 3 12 23 13

12

3

J

(4-25)

In the above equations, both expressions in terms of strain and principal

values are given. Upon further examination, the second invariant of plastic strain

Page 115: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

115

is non-negative while the third invariant of plastic strain has the negative sign

function of the intermediate plastic strain, appropriate for the sign-sensitive

Bauschinger effect. Additionally, the second and third invariants of plastic strain

are suitable for describing asymmetric isotropic hardening because they are

constant in the elastic regime and continuous in the elastic-plastic.

4.2.1 New Description of Hardening

With a broader definition of isotropic hardening to include distortion of the

yield surface that does not annihilate the material isotropy, it is necessary to

examine the evolution of the yield surface under different load and strain paths.

Note that the evolution of tension-compression asymmetry can be directly

observed under reversed loading or inferred from the increment of plastic strain

in torsion testing of thin-walled cylinders using the associated flow rule. In

particular, the evolution of the asymmetry of ES-1 was evaluated using the cyclic

test data presented in Section 3.4.

For such cyclic loading, first ES-1 material was subject to tensile strain

and presumably uniaxial tensile stress, i.e.:

0 0

0 0 0

0 0 0

T

σ (4-26)

20 0

3

10 0

3

10 0

3

T

s . (4-27)

Under this state of stress, the second and third invariants are:

Page 116: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

116

22 2 2

2 1 2 3

3

3 1 2 3

1

2 3

2

27

T

T

J s s s

J s s s

. (4-28)

Whereas the strain is:

1 0 0

10 0

2

10 0

2

T

ε (4-29)

22 2 2

2 1 2 3

3

3 1 2 3

31

2 4

4

T

T

J

J

. (4-30)

Initial yielding was observed as the specimen started to deform plastically.

At maximum displacement (0.57 mm), the flow stress was recorded and the

motion was reversed relieving the uniaxial tensile stress and then creating a state

of uniaxial compressive stress in the elongated specimen (assuming no bending

or buckling within the specimen gage length). In this stage of the test,

0 0

0 0 0

0 0 0

C

σ (4-31)

and its deviator is:

20 0

3

10 0

3

10 0

3

C

s (4-32)

Page 117: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

117

while the second and third invariants are:

22 2 2

2 1 2 3

3

3 1 2 3

1

2 3

2

27

C

C

J s s s

J s s s

. (4-33)

Still under compressive stress, the specimen was returned to its initial

configuration 2 3 0J J and subject to compressive strain. Again, yielding

was observed as the specimen began to plastically deform, this time in

compression. The process was repeated 50 times in accordance with standard

test procedures.

1 0 0

10 0

2

10 0

2

C

ε (4-34)

22 2 2

2 1 2 3

3

3 1 2 3

31

2 4

4

C

C

J

J

(4-35)

To minimize the kinematic softening of the cyclic response, the first cycle

at each value of plastic strain of the cyclic test data of forged ES-1 in Figure 4-3

was used to find the limits of the elastic region after two load reversals between

uniaxial tension and compression.

Page 118: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

118

Figure 4-3. The first stress-strain cycle of forged ES-1 during cyclic testing at completely reversed displacement of 0.57, 0.62 and 0.67 mm and the uniaxial tensile and compressive flow stress.

The offset, between zero (i.e. the proportionality limit) and 0.2% (the

ASTM standard), used to determine yielding following the load reversal greatly

affected the calculation of the flow stress and thus the evolution of asymmetry.

Still, the maximum absolute stress and 0.1% offset were used to determine the

limits of the elastic region in Figure 4-4.

σT(ε

T

P)

σT(ε

C

P)

σC(ε

C

P)

σC(ε

T

P)

Page 119: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

119

Figure 4-4. Flow stress at the limits of the elastic region by plastic strain in forged ES-1 under cyclic loading.

As shown in Figure 4-4, the limits of the elastic region are increasingly

asymmetric with accumulation of plastic strain, a phenomenon known as the

Bauschinger effect. In kinematic hardening models, the asymmetry is commonly

modeled as a translation of the yield surface due to a fictitious back stress. The

back stress according to the most common kinematic hardening models

presented in Chapter 1 are functions of the increment of plastic strain tensor.

Therefore, under uniaxial tension, the back stress is also uniaxial. The uniaxial

back stress at various levels of plastic strain is shown in Figure 4-5.

1 1

3 3

1 1

3 3

3 3 2 3 3 2

3 3 2 3 3 2

T Cc c

c c

(4-36)

Page 120: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

120

Figure 4-5. The uniaxial back stress as a function of plastic strain for forged ES-1 under cyclic loading.

An example of the translation of the yield surface using a kinematic

hardening law is presented alongside the classical isotropic hardening surface

and the asymmetric hardening surface for comparison in Figure 4-6. This shows

that there are two major disadvantages to these kinematic hardening laws. First,

the back stress is non-zero even after unloading (i.e. after all stresses have been

removed from the material). Secondly, a large disparity between the monotonic

and kinematic yield surfaces especially pronounced along the third compressive

meridian is evident upon initial loading (Figure 4-6). Additionally, there are also

limitations to the kinematic hardening models. The translation destroys the

isotropy, which is inappropriate for materials that remain isotropic even after such

loadings.

Page 121: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

121

Figure 4-6. Subsequent yield surfaces using classical isotropic, redefined isotropic, and kinematic hardening under uniaxial tensile loading.

Another possible explanation for increasingly asymmetric flow stresses at

the limits of the elastic region under cyclic loading is that the evolution of

asymmetry depends on the strain tensor or some component thereof not just the

equivalent plastic strain. This possibility is attractive because a separate

kinematic hardening may no longer be required to model load reversals, the yield

surface remains symmetric about zero stress for isotropic materials, and the new

hardening law allows description of tension-compression asymmetry. To further

explore this idea, the asymmetry parameter was calculated assuming that the

yield surface did not translate. The evolution of the asymmetry parameter, c,

Page 122: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

122

which was calculated at each level of plastic strain using Eq. (4-8), is included in

Figure 4-7.

Figure 4-7. Evolution of the material asymmetry parameter, c, for forged ES-1 using the limits of the elastic region under cyclic loading.

An innovative evolution law was proposed to describe the data. Namely,

the evolution of c was modeled assuming that it depends on the third invariant of

plastic strain. It was approximated:

30

1

tanhJ

c cc

. (4-37)

where the parameters of ES-1 are 0 1.7c and 1 0.002c . Note the asymmetry

parameter approaches c0 as the third invariant approaches infinity, and negative

c0 as it approaches negative infinity. The uniaxial flow stresses and superposed

corresponding yield surface evolution in the deviatoric plane for several values of

plastic strain are included in Figure 4-8. Note that under tensile (compressive)

plastic strain the yield surface hardens near the tensile (compressive) meridians

and softens near the compressive (tensile) meridians. Also notice that only

Page 123: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

123

dependence of c on the third invariant of plastic strain was considered, but the

effects of the second invariant of plastic strain could also be of interest.

Additionally, it is worth mentioning that the elastic-plastic model with asymmetric

isotropic hardening results in stabilizing hysteresis loops under cyclic loadings.

This is in contrast with the classical isotropic hardening which cannot reproduce

such effects.

Figure 4-8. Uniaxial flow stress at the limits of the elastic region and yield surfaces in the deviatoric plane for several values of plastic strain in forged ES-1 under cyclic loading.

It is worth noting the difficulties associated with correct estimate of the

evolution of yield surfaces based on cyclic tests. For example, Paul (1968)

discusses three acceptable methods of determining the yield stress: Lode

extrapolation method, plastic strain offset, and the proportionality limit technique.

The method used for identifying the yield stress may create large differences in

the values of the asymmetry parameter, c, and, therefore, on the evolution of the

yield surface. Hence, it is proposed to obtain the evolution of c using pure shear

Page 124: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

124

deformation data as described in the next section. However, the approach used

so far may be more realistic in modeling cyclic data through kinematic hardening.

4.2.2 Distortional Hardening with Asymmetry

The idea for the different method for determining the asymmetry

parameter sprang from the necessity to model the Bauschinger effect frequently

associated with kinematic hardening. However, the concept can be used to

better model monotonic loading as well. Since the asymmetry parameter is

dependent on the second and third invariants of plastic strain, the subsequent

yield surfaces will differ under monotonic uniaxial tension, monotonic uniaxial

compression, and pure shear strain as shown in Figure 4-9. These three

different experimental characterizations can be used to parameterize the yield

function with asymmetric hardening.

Page 125: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

125

Figure 4-9. Cazacu and Barlat subsequent yield surface in the deviatoric plane following uniaxial tension, compression, or pure shear strain.

Let us recall that under monotonic pure shear strain, when the third

invariant of plastic strain is equal to zero, ultra high-strength steels strain harden

according to a power-type law as a function of shear strain (which, in this case, is

equal to the second invariant of plastic strain). Thus in general, the evolution of

the uniaxial yield stress and that of the uniaxial compressive yield stress can be

described:

2 3 0 2 2 3 3

2 3 0 2 2 3 3

,

,

T T T T

C C C C

J J J J

J J J J

, (4-38)

Page 126: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

126

Where 2 3,T J J and 2 3,C J J are the tensile and compressive flow stresses

following uniaxial tensile and compressive stress, 0T and

0C are the initial

tensile and compressive yield stresses, 2 2T J and 2 2C J are the

components of tensile and compressive hardening resulting from the second the

invariant of plastic strain, and 3 3T J and 3 3C J are the components of

tensile and compressive hardening resulting from the third invariant of plastic

strain, respectively.

Similarly, for the remainder of the yield surface, the total hardening is a

combination of hardening as a function of the second and third invariants of

plastic strain. The decomposed elastic regions resulting from monotonic uniaxial

tensile loading are shown in Figure 4-10 alongside the subsequent yield surface.

The initial and subsequent yield surfaces are depicted together to illustrate the

material softening for regions of the yield surface near the compressive

meridians. The components are discussed in the following sections.

Figure 4-10. Decomposition of the yield surface with distortional hardening A) initial yield surface, B) second invariant component of the

subsequent yield surface, C) third invariant component of the subsequent yield surface, and D) the total subsequent yield surface.

3T

sT 0

3 2'J J

sT 2

sT 3

sT

Initial Yield Surface J2

ε Component of the

Yield Surface

J3

ε Component of the

Yield Surface

Subsequent Yield Surface

2 2'J J 3 2'J J

sz(MPa) s

z(MPa) s

z(MPa) s

z(MPa)

sy(MPa) s

y(MPa) s

y(MPa) s

y(MPa)

sC 2

2 2'J J

Page 127: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

127

4.2.2.1 2J -component of the hardening law

As seen in Figure 4-10, the second invariant of total plastic strain is

associated primarily with a growth of the subsequent yield surface in Haigh-

Westergaard space and can be fully understood by a single fixed-end torsion test

or another test under pure shear strain for example using a butterfly specimen.

Based on the symmetry of the second invariant of total plastic strain, the 2J -

component of the subsequent yield surface is the same under uniaxial tensile

and compressive loading as shown in Figure 4-11. The uniaxial tensile and

compressive hardening, however, need not be equal (i.e. the model allows for

evolution of asymmetry under pure shear strain). For example, UHSS with

material tendency toward greater strength in compression than tension based on

the BCT microstructure, also results in a slight change in shape of the

subsequent yield surface (i.e. the increment of hardening associated with the

second invariant of plastic strain is somewhat greater along the compressive

meridians than the tensile meridians). This evolution of asymmetry with

accumulated plastic strain is a material property, c2, that can also be determined

from a single pure shear strain test using the associated flow rule. The

parameterization of the material asymmetry parameter, c2, and then the

hardening, Y, are discussed in the next subsection.

Page 128: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

128

Figure 4-11. Cazacu and Barlat yield surface under both uniaxial tension and compression and the common component.

4.2.2.1.1 Material asymmetry parameter

It is assumed that pure shear deformation 3 0J can be produced by a

combination of shear and axial loading. There are six points on the yield surface

such that the strain increment is parallel to a shear axes and perpendicular to a

meridian as shown in Figure 4-12.

Page 129: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

129

Figure 4-12. Initial and subsequent Cazacu and Barlat yield surface under pure shear strain.

Using the point on the yield surface in the 3-plane where 3

0

, the

strain, invariants of strain, state of stress, deviatoric stress, and the deviatoric

stress invariants are presented below.

0 0 0 0

0 0 0 0

0 0 0 0 0 0

ε (4-39)

2 1 2 3

3 1 2 3

1

2

0

J

J

(4-40)

0

0 0

0 0 0

σ (4-41)

J2

ε Component of the

Yield Surface

J '2

et

P( ) 3 2'J J

sT 2

sz(MPa)

sy(MPa)

sC 2

f1 f

2

f3

τ12

τ31

τ23

Page 130: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

130

22

22

0 06 4

0 06 4

0 03

s (4-42)

22 2 2 2

2 1 2 3

3 2

3 1 2 3

1'

2 3

2'

27 3

J s s s

J s s s

(4-43)

The (‘) indicates the second and third invariants of deviatoric stress under

pure shear deformation.

Under pure shear deformation in the 3-plane under the described loading,

the evolution of asymmetry as a function of shear strain is calculated based on

the associated flow rule for the instantaneous yield surface where the through

plane strain is zero.

Page 131: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

131

1 2 2 2 22 22

2

231 322 3

1 2 2 2 22 22

2

232 322 3

1 2 2

22

233 322 3

3'

2 6 4 18 3 4 3

3 ' '

3'

2 6 4 18 3 4 3

3 ' '

2'

2 9 3

3 ' '

J c

J cJ

J c

J cJ

J c

J cJ

0

(4-44)

Solving the third partial derivative for the material asymmetry parameter,

c2:

22

2 2 2

93

2 6c

. (4-45)

Using the Cazacu and Barlat (2004) yield criterion, uniaxial tensile and

compressive flow stresses following pure shear deformation as a function of the

second invariant of plastic strain are on the same subsequent yield surface with

the shear flow stress. Thus, the hardening of the uniaxial tensile and

compressive flow stresses following pure shear strain are:

Page 132: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

132

13 3

1 2 3 223 23

222 2 3

2 2 0 01 1

3 32 2

13 3

1 2 3 223 23

222 2 3

2 2 0 01 1

3 32 2

23

3 ' ' 3 27 3

3 3 2 3 3 2

23

3 ' ' 3 27 3

3 3 2 3 3 2

T T T

C C C

cJ c J

J

c c

cJ c J

J

c c

(4-46)

4.2.2.1.2 Hardening under pure shear strain

The hardening law that relates the yield surface in the six dimensional

stress and strain spaces describes the six points of pure shear deformation (i.e.

3 0J ) common to all subsequent yield surfaces of the material.

1 13 33 32 22 3 2 3' ' 0J cJ J cJ

(4-47)

Note that the asymmetry parameter is not the material asymmetry

parameter, c2, but the asymmetry parameter as a function of the second and

third invariants of plastic strain defined in the following section. Equation (4-47)

means that all possible subsequent yield surfaces for any given value of

equivalent plastic strain, intersect at the six points where 3 0J (Figure 4-9).

Distortional hardening under general three-dimensional loads will result in the

same hardening behavior of the six points of pure shear deformation, but the

effective stress and plastic strain increments will vary. It is more conventional to

express the hardening as the shear stress in pure shear strain as a function of

the equivalent plastic strain.

Page 133: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

133

n

P

Y k

(4-48)

In order to express the hardening as a function of equivalent plastic strain,

the axial stress is expressed as a function of the material asymmetry parameter

and the shear stress using Eq. (4-45). The fourth order polynomial is:

2 4 2 2 2 2 4

2 2 24 27 48 81 144 0c c c (4-49)

By solving the quadratic equation, the following equation is obtained:

2 2

2 22 2

2

2

3 27 16 9 3 32 27

2 4 27

c c

c

(4-50)

It is convenient to use the axial-shear stress ratio, r, defined:

1

22 2

2 2

2

2

3 27 16 9 3 32 27

2 4 27

c cr

c

(4-51)

The effective stress under pure shear strain can then be expressed as a

function of the axial-shear stress ratio.

11 3 3

3 3 22 32

2 3

1 2 1' ' 1

3 27 3J cJ r c r r

(4-52)

Finally, substituting the effective stress in pure shear strain into Eq. (4-47)

and expressing the hardening as a shear strain power law, the yield function is:

13 322 3

13 32

2 3

0

1 2 11

3 27 3

nP

J cJ

k

r c r r

(4-53)

Page 134: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

134

The equivalent plastic strain is usually defined by using the work

equivalency principle based on the assumption that an equal amount of work is

required to harden the material to the same subsequent yield surface regardless

of the stress and plastic strain tensors. Since this proposed distortional model

results in a different subsequent yield surface for each state of plastic strain,

consistency lies in the six points of shear hardening. Therefore, the equivalent

plastic strain for monotonic loading is based only on the second invariant of strain

similar to the von Mises equivalent plastic strain. Recall, however, that even this

von Mises expression of equivalent plastic strain is only valid under monotonic

loading. Difficulty arises during strain reversals when the magnitude of total

plastic strain and, necessarily, the second invariant of plastic strain are reduced

because the flow stress at the point of loading should never decrease. The

hardening being based on the idea that compressive strain results in

compressive strain hardening and tensile strain produces tensile strain

hardening.

Nonetheless, based on monotonic loading, the equivalent plastic strain is

dependent on the second invariant. For this research, the scalar plastic strain is

described for monotonic loading as the square root of the second invariant of

plastic strain, similar to von Mises equivalent plastic strain.

2

1

222

2

1

2 2

P

P

P P

J

JJ

J

ε (4-54)

Page 135: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

135

It is necessary that the increment of equivalent plastic strain be non-

negative. Therefore, the product of the absolute values of the derivative of

equivalent plastic strain and the increment of plastic strain are used in Eq. (4-56).

0Pd (4-55)

22

P

P Pd dJ

ε

ε (4-56)

Now, with the six points of shear strain and the material asymmetry, the

3J -component of hardening and the Bauschinger effect will be discussed in the

subsequent section.

4.2.2.2 3J -component of the hardening law

The third invariant of plastic strain then is responsible for both the

Bauschinger effect under reversed loading and the increasing asymmetry in the

yield surface under combined tension torsion loading determined using the

associated flow rule as shown in Figure 4-13. Since there is not a state of strain

with a nonzero 3rd invariant of plastic strain and a zero second invariant of plastic

strain, the other 3J -dependent component can be identified by subtracting the

first component from the total elastic region with a known contribution of the

second and third invariant of plastic strain. For simplicity, uniaxial tension and

uniaxial compression tests can be used to parameterize the asymmetry

parameter for positive and negative 3rd invariants of plastic strain, respectively.

Page 136: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

136

Figure 4-13. Cazacu and Barlat initial and third invariant component of the subsequent yield surfaces using asymmetric isotropic hardening under uniaxial tensile loading in the deviatoric plane.

Notice the 3J -component of the subsequent yield surface under uniaxial

tensile loading in Figure 4-13 shows hardening near the tensile meridians and

softening near the compressive meridians, the Bauschinger effect. Furthermore,

the surface remains isotropic, centered at the origin and maintaining three-fold

symmetry about the projections of the principal axes.

Next, to parameterize the evolution of asymmetry under axial strains, the

component of the uniaxial tensile flow stress associated with the third total plastic

strain invariant is determined by linear decomposition.

3 3 2 2 0

3 3 2 2 0

P

T T T T

P

C C C C

J J

J J

(4-57)

sy(MPa)

sz(MPa)

sT 3

Page 137: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

137

Note that the tensile flow stress is characterized for both positive and

negative values of the third invariant of plastic strain using the uniaxial

compression test data. Now, the asymmetry parameter is determined as a

function of the uniaxial tensile flow stress which is a function of the second and

third invariants of total plastic strain.

33

2 2 32322

3 2 2 3

3

3 9 3 327 ' 3 3

27 ' 2 2 9 2

TT

T T

Jc

J

(4-58)

4.3 Implementation of the Distortional Hardening Model For illustrative purposes, ES-1 was first modeled using the Cazacu and

Barlat (2004) yield criterion with conventional isotropic hardening and then, for

comparison, using the von Mises yield criterion with the proposed distortional

hardening model.

It was assumed that ES-1 was a Cazacu and Barlat (2004) material at the

outset of this research. Therefore, using the flow stress of Eglin steel in tension

and compression, the material constant (c) was determined at several values of

plastic strain to characterize the progression of the tension-compression

asymmetry. The evolution of the material parameter, c, with the accumulated

plastic strain is shown in Figure 4-14.

Page 138: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

138

Figure 4-14. Cazacu and Barlat material parameter, c, by plastic strain for forged specimens, cast and HIP’d specimens and cast specimens.

The yield surface was plotted in the deviatoric (y-z) plane using the

constant, c, and the normalized stress in uniaxial tension as shown in Figure

4-15.

Page 139: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

139

Figure 4-15. Cazacu and Barlat yield surfaces in the deviatoric plane (A) forged, (B) cast and HIP’d, and (C) cast no HIP specimens.

Page 140: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

140

If only dependence of P is considered, variation of c can be

approximated as:

ln P Pc A B C (4-59)

where the constants A, B, and C were determined using the experimental data

and are presented in Table 4-1.

Table 4-1. Cazacu and Barlat yield criterion parameters. Coefficients Mises Cazacu Constant Linear Cazacu Log Cazacu

A 0.0 0.000 0.000 -0.166 B 0.0 0.000 -13.170 0.000 C 0.0 -0.226 0.047 -0.966 A 0.0 0.000 0.000 -0.269 B 0.0 0.000 -21.220 0.000 C 0.0 -0.124 0.315 -1.318 A 0.0 0.000 0.000 -0.208 B 0.0 0.000 -13.60 0.000 C 0.0 -0.161 0.183 -1.216

In order to more accurately determine the value of c at much higher values

of local plastic strain, the linear interpolation procedure was used (see Table

4-2).

Page 141: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

141

Table 4-2. Points for linear interpolation of the Cazacu and Barlat asymmetry parameter, c.

Equivalent Plastic Strain Forged Cast and HIP’d Cast no HIP

0.00 -0.2120 0.0000 -0.0682 0.02 -0.3119 -0.2079 -0.1335 0.04 -0.3387 -0.3152 -0.1511 0.06 -0.3548 -0.3792 -0.1617 0.08 -0.3663 -0.4250 -0.1693 0.10 -0.3753 -0.4605 -0.1752 0.12 -0.3827 -0.4895 -0.1800 0.14 -0.3889 -0.5140 -0.1842 0.16 -0.3944 -0.5352 -0.1877 0.18 -0.3992 -0.5538 -0.1909 0.20 -0.4034 -0.5705 -0.1934 0.22 -0.4073 -0.5855 -0.1963 0.24 -0.4109 -0.5992 -0.1986 0.26 -0.4141 -0.6118 -0.2008 0.28 -0.4171 -0.6234 -0.2028 0.30 -0.4199 -0.6342 -0.2046 0.32 -0.4226 -0.6443 -0.2064 0.34 -0.4240 -0.6538 -0.2080 0.36 -0.4274 -0.6627 -0.2095 0.38 -0.4296 -0.6711 -0.2110 0.40 -0.4317 -0.6791 -0.2124

This distortional hardening model in conjunction with a von Mises yield

criterion reveals very interesting trends. Recall that the initial yield surface for a

Mises material is circular in the deviatoric plane. Under pure shear deformation

as in a fixed-end torsion test, the proposed hardening model predicts that the

circle will grow in size, but retain the same circular shape as shown in Figure

4-16. However, under axial loading 3 0J , the yield surface would begin to

take the form of a rounded triangle. Under tension, the surface would have

maximum curvature at the intersection of the three tensile meridians and reach a

minimum curvature at the intersection with the compressive meridians.

Conversely, under compression, the surface would reach maximum curvature at

the intersection with the three compressive meridians and reach a minimum

curvature at the intersection with the tensile meridians.

Page 142: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

142

Figure 4-16. Cazacu and Barlat yield surface in the deviatoric plane under uniaxial tension, uniaxial compression, and pure shear using distortional hardening.

Furthermore, for a von Mises material with such distortional hardening, the

evolution of asymmetry is independent of the second invariant of plastic strain

2 0c . Therefore the yield function reduces to the familiar form.

13 322 3 YJ cJ

(4-60)

where c is an odd function of the third invariant of plastic strain. For this model,

the asymmetry parameter was approximated again using the hyperbolic tangent

Page 143: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

143

function from the cyclic characterization and power law hardening in pure shear

strain.

31.7 tanh0.002

Jc

(4-61)

4.4 Implications of the Proposed Distortional Hardening Model

While the idea for this new model for hardening came from the need to

better account for the Bauschinger effect, induced tension-compression

asymmetry, and how it affects the intrinsic asymmetry due to the structure of the

material, the new model has broader applicability. Let us point to its advantages.

First, as three different monotonic stress-strain curves are used to parameterize

the subsequent yield surfaces, the model is accurate under monotonic uniaxial

tension, uniaxial compression, and shear strain. Second, the model captures the

trend of induced tension-compression asymmetry under reversed loading often

seen in dynamic or cyclic applications or the effect of pre-strained components.

Third, the model predicts the direction of the plastic flow under general three-

dimensional loads.

Expressly under dynamic conditions, reversals of three-dimensional

loadings are possible as waves propagate through the material. With accuracy in

uniaxial tension, uniaxial compression, and pure shear plastic strain under

monotonic loading as well as capturing the trend of softening under reversed

loading, it may be possible to provide the most accurate predictions under three-

dimensional, dynamic conditions using this relatively simple yield function.

This model provides new insights and paths to be taken for further

understanding of asymmetric hardening. Much more research is required to

Page 144: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

144

understand the effects of strain rate, temperature, and orthotropy on the evolving

yield surface. Additionally, more research is required to glean more predictive

capability from the model. For example, is the extent of asymmetry predictive of

aging, spring-back, strain localization, or failure?

Creep and high strain-rate tension tests could reveal the rate-sensitivity of

each component of the asymmetric hardening laws.

Finally, while this research concerns isotropic behavior, the applicability of

this new hardening model to anisotropic materials is a promising research path.

The theoretical yield function parameterized using experimental data was

then implemented in a numerical solution in a user material subroutine within

ABAQUS implicit. The finite element analysis is discussed in Chapter 5.

Page 145: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

145

CHAPTER 5 FINITE ELEMENT ANALYSIS OF EGLIN STEEL

The capabilities of the model developed for Eglin steel to capture the

mechanical behavior for general loadings will be assessed. To solve boundary-

value problems for plastically deformable materials, generally the finite element

framework is used. Therefore, first a brief overview of this framework including

the strong and weak form of the balance of linear momentum and finite element

discretization is presented along with the user material subroutine (UMAT)

containing a time-integration algorithm used for implementing the model in

commercial FEA software (ABAQUS, 2009). Next, finite element simulations of

the mechanical response of the material for a variety of loading conditions for

both quasi-static and dynamic loadings are presented. Comparison between

simulation results and data attest to the adequacy of the formulation proposed.

Moreover, the model is used to gain insights into the behavior of ES-1 for

loadings that are not accessible to direct experimentations. Such numerical tests

are useful to engineers in the field who need to assess the behavior of the

material for combined loadings involving shear.

5.1 Finite Element Formulation

The implicit solver of the commercial software ABAQUS (see ABAQUS,

2009) was used in the finite element analysis. Let Φ ,tX define the motion of a

deformable material from the reference to the current configuration. This

software uses an updated Lagrangian formulation, that is Φ ,t+ t at the

increment n+1 corresponding to the instant t+ t is calculated based on Φ ,t at

Page 146: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

146

the increment n at the instant t by imposing the velocity field. Thus, the

deformation gradient, Φ

XF

, is:

,t t

iij t

j

XF

X

(5-1)

and all the dependent variables i.e. the Cauchy stress σ , the velocity, v, and the

strain-rate tensor D are expressed as functions of the material coordinates X.

Let be the domain occupied by the plastically-deformable material with

the boundary at time t. For isothermal processes, the governing equations

include the balance of linear momentum and the constitutive equations relating

the stress to the rate of deformation or strain-rate, D. Boundary conditions, such

as tractions T on t and velocity v on v , with t v , are associated

with the governing equations. The strong form consists of the balance of the

linear momentum presented here with zero body forces and traction boundary

conditions:

div σ a on

T n σ T on (5-2)

Where is the current density, and a is the acceleration (the second derivative

with respect to time of the displacement vector, =a u ). The above equations

depend only on the velocity field v, because the stresses can be expressed in

terms of velocities by the constitutive model. The exact solution to this set of

differential equations is difficult, if not impossible, to calculate. Hence, to

approximate the solution by discretization of the domain into finite elements with

n nodes, the strong form is multiplied by a vector-valued weighting function ( v )

Page 147: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

147

that is zero on the boundary v and then integrated over the volume. Note that

since stress is symmetric:

: :v σ σ D

where 1

2

Tv v D = , the weak form can be written as:

:a vd vd T vdA

. (5-3)

An admissible trial solution for the element velocity ev is differentiable

and satisfies the boundary conditions (i.e. e vv on

v ). The acceleration ea is

given by the time derivative of the trial function.

1

,n

e k k

k

t N t

v X X v , (5-4)

1

,n

e k k

k

t N t

a X X a , (5-5)

where kN ( )X denote the shape functions having the desired level of continuity

such that ev is admissible. Using the Galerkin method, the weighting function

has the same form as the trial solution with the additional restriction v = 0 on

v .

1

n

k k

k

v N v t

X X (5-6)

The rate of deformation is the symmetric part of the spatial velocity

gradient (i.e. derivatives are taken with respect to the spatial coordinate x and

not X):

Page 148: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

148

ij Ii I, j Ij I,i

1

1D = v N v N

2

n

ek

.

Substituting the weighting and trial functions back into the weak form, the

equation of motion is assembled into a global system of equations,

: 0

e e e

I e e I e I eN d N d N dA

a σ T (5-7)

Or in matrix form:

ext intM [a]= f f (5-8)

In the above equation the matrix M is defined as

e

IJ I JM N N

, (5-9)

while f dA

et

ext

I I eN

T and f dΩ

et

int

k k eN

. For more information on finite

element analysis the reader is referred to the textbook by Belytschko et al.

(2013).

Moreover, it is important to note that in ABAQUS, all stresses and strains

are rotated by R, the proper-orthogonal rotation corresponding to the polar

decomposition of the gradient of deformation (i.e. F = RU = VR) with the right

stretch tensor (U) and left stretch tensor (V) being symmetric and positive-definite

tensors before the UMAT is called.

ABAQUS uses of the Green-Naghdi rate (Green and Naghdi, 1965) which

is objective (frame invariant) and has the remarkable property that it reduces to a

time derivative in the rotated coordinate system. Thus, in ABAQUS, the finite-

deformation formulation of any elastic-plastic model has the same form as the

Page 149: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

149

small deformation formulation because D reduces to ε in the rotated frame. For

proofs on the unified treatment of finite-deformation and small-deformation

theories, the reader is referred to Hughes (1984).

Without a UMAT, an elastic-plastic model using the von Mises (1913) yield

criterion is used to solve the global system of equations in ABAQUS implicit. The

UMAT described in the following section was developed to implement the

viscoplastic model with the Cazacu and Barlat (2004) yield criterion and the

proposed distortional hardening model in the finite element analyses.

5.2 Implementation of the Elastic-Plastic Model in FEA Recall in this research, the yield function is expressed as the difference

between the effective stress (a function of the Cauchy stress and the plastic

strain) and the hardening (a function of the plastic strain).

, ,P P PF Y σ ε σ ε ε (5-10)

The total plastic strain increment is commonly expressed as a summation

of the plastic strain increment and the elastic strain increment using Hooke’s law.

1

:P e P eC

(5-11)

Equations (5-10) and (5-11) are combined and solved using the iterative

Newton-Raphson numerical root-finding algorithm in the UMAT.

1 1 1 1 1

1

1 1 1

, , 0

: 0

P P P

n n n n n n

e P

n n n

F Y

σ ε σ σ ε ε

ε C σ ε (5-12)

At each increment of time 0 through n, the above system of equations is

solved using an iterative root finding algorithm called the Newton-Raphson

method using a Taylor series expansion within the UMAT.

Page 150: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

150

1 1 1 11 1

1 1 1 1

1 11 1

1 1 1 1 1

, ,, : : 0

: : 0

m Pm m Pm

n n n nm P m m Pm

n n n nP

e m Pm e m Pm

n n n n n

F FF

σ ε σ εσ ε σ ε

ε C σ ε C σ ε

(5-13)

For the first iteration of the Newton-Raphson, the initial trial solution (m=0)

is assumed to be completely elastic. Hence, the trial solution for the stress

tensor is described by Hooke’s law and the plastic strain increment is zero.

1 1:trial e

n n n σ σ C ε (5-14)

If the yield function was not within the specified tolerance (usually 1 Pa),

the trial solution was updated (m+1) by a variation, 1

1

m

n

σ and 1

1

Pm

n

ε .

1 1

1 1 1

1 1

1 1 1

Pm Pm Pm

n n n

m m m

n n n

σ σ σ (5-15)

Recall that the increment of plastic strain is normal to the yield surface.

21 1 1

1 1 1 121

:

m

pm m m m

n n n nn

F F

ε σ

σ σ

Solving the system of equations for the variation in the stress:

11 1

1 1 1 1 11 1

: :

m m

m m m m

n n n n npn n

F FF

eσ P C σ εε σ

(5-16)

1

p

F F

eP C

σ ε (5-17)

11 1

1 1 1 1:Pm m Pm

n n n n

eε ε C σ ε (5-18)

This technique is better known as return mapping. Once the yield function

is equal to zero within a specified criterion, 1 Pa, the process is repeated for the

next time increment. The stress and plastic strain are updated until the

Page 151: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

151

numerical analysis converges upon a solution at each increment of time. The

UMAT was used in the finite element analyses discussed in Section 5.3.

5.3 Finite Element Analysis of Ultra High Strength Martensitic Steel

First, verification of the implementation of the model using a right

rectangular prism was completed. Next, all of the experimental loading scenarios

were simulated and compared with the experimental data. Let us recall that the

flat specimens were used to characterize the uniaxial tensile response for model

identification. Validation of the model was performed by comparing the

simulation results with data from tests on round specimens. In addition, the

capabilities of the model to predict the shear response of ES-1 were investigated.

Finally, the model with a rate-dependent hardening rule was used to predict the

behavior for dynamic conditions. A summary of all the finite element simulations

is included in Table 5-1.

Page 152: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

152

Table 5-1. Finite element analysis simulation matrix.

Strain Rate Manufacturing Process

Loading Configuration

Specimen Shape

Quasi-Static Forged Tension Round

Pin Loaded

Prism

Compression Round

Torsion Thin Walled

Cast and HIPd Tension Round

Pin Loaded

Compression Round

Torsion Thin Walled

Cast Tension Round

Pin Loaded

Compression Round

Torsion Thin Walled Dynamic Forged SHPB Round

Taylor Impact Round Cast and HIPd SHPB Round

Taylor Impact Round

Cast SHPB Round

Taylor Impact Round

5.3.1 Verification of the FE Implementation

The finite element analysis of a right rectangular prism (1 mm x 1 mm x 2

mm, see Figure 5-1) composed of two linear hexahedral elements with reduced

integration (C3D8R) subject to uniaxial tension was used to verify the UMAT

development. Symmetry and displacement boundary conditions were

prescribed.

Page 153: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

153

Figure 5-1. Stress distribution according to the model within forged ES-1 under

quasi-static uniaxial tension.

Next, the axial force require to obtain the prescribed displacement was

used to calculate the stress. It was then verified that this stress-strain response

matches the input hardening law (Swift law) as well as the stress-strain curve

obtained directly at the integration point of one element (see Figure 5-2) for

comparison of the three curves. Note the implementation is correct with slight

differences associated with elastic contributions.

Page 154: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

154

Figure 5-2. Stress and plastic strain for forged ES-1.

5.3.2 Comparison between Model Predictions and Experimental Results in Uniaxial Tension and Compression

The proposed model was further applied to simulate the plastic response

of the flat tensile specimen (for the geometry of the specimens, see Chapter 3)

and compared with the experimental observations acquired using DIC. The finite

element mesh for the flat tensile specimens consisted of 2829 linear hexahedral

element with reduced integration (C3D8R).

For each processing condition (i.e. forged, cast, and cast and HIP’d), the

finite element predictions of the strain isocontours were compared to the DIC

local strain maps obtained experimentally for the same axial displacement, the

onset of necking.

Page 155: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

155

Figure 5-3. The longitudinal strain distribution from DIC and FEA forged specimens.

Figure 5-4. The longitudinal strain distribution from DIC and FEA cast and HIP’d specimens.

Page 156: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

156

Figure 5-5. The longitudinal strain distribution from DIC and FEA of cast specimens.

Note that there is an overall agreement between model and data for all

materials. Next, the predictions of the global behavior in uniaxial tension were

compared to the experimental data for all materials. Note the excellent

agreement in Figure 5-6.

Page 157: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

157

Figure 5-6. Stress-strain curves of flat DIC and FEA specimens in tensile loading A) forged, B) cast and HIP’d, and C) cast.

Page 158: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

158

Next, the finite element predictions were compared to data for each

material under uniaxial compression. The finite element mesh of 3375

hexahedral elements is presented in Figure 5-7 and the specimen dimensions

are included in Chapter 3. Note the excellent agreement obtained for all

materials for this loading in Figure 5-8.

Figure 5-7. The uniform, longitudinal stress distribution within the forged round finite element specimens under quasi-static compressive loading

A) cross section and B) side.

Page 159: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

159

Figure 5-8. Stress-strain curves of round DIC and FEA compression specimens A) forged, B) cast and HIP’d, and C) cast.

Page 160: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

160

5.3.3 Model Validation: Comparison with Uniaxial Tension in Axisymmetric Specimens

In order to further assess the predictive capabilities of the plasticity model,

the model predictions have been compared to the uniaxial tension experiments

performed on axisymmetric specimens. The geometry of the specimen is given

in Chapter 3. Due to symmetry of the problem only an eighth of the specimen

was modeled using 6400 hexahedral elements with reduced integration. The

mesh was refined in the mid cross-section region as shown in Figure 1-11 in

order to capture the localization of the deformation. Boundary conditions were

applied such as to maintain the symmetry of the problem, and an axial

displacement of 3mm was imposed at the extremity of the specimen.

For the (A) forged, (B) cast and HIP’d, and (C) cast specimen, the

predicted load-displacement curves are compared with the experimental ones

(see Figure 5-9). It is worth mentioning the agreement between the model

prediction and the experiments. Also, in Figure 5-9, it could be seen that

numerical predictions considering a Swift law as an isotropic hardening law are

more accurate than the one using a Voce hardening law. Furthermore, the

discrepancies observed between the predicted load-displacement curve for the

elastic domain are due to the fact that the experimental displacement has been

acquired using the crosshead, i.e. the measured displacement encompass the

compliance of the load frame.

To overcome this issue, the experimental displacement has also been

accurately measured using the DIC technique. A virtual extensometer has been

placed in the gauge length of the specimen to precisely record the axial strain. In

Page 161: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

161

Figure 5-10 is shown the comparison between the predicted stress-strain curve

and the experimental one with the strain calculated using DIC measurements. It

is worth noticing the very good agreement between the model prediction and the

experiments for the three materials.

Page 162: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

162

Figure 5-9. The experimental and model load-displacement curves A) forged, B) cast and HIP’d, and C) cast materials.

Page 163: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

163

Figure 5-10. Stress-strain curves for round DIC and FEA tension specimens A) forged, B) cast and HIP’d, and C) cast.

Page 164: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

164

As an example, Figure 5-11 shows the predicted isocontour of the von

Mises stress for the forged ES-1 material at different axial displacements

corresponding to the occurrence of plastic strain (axial plastic strain of 0.2%,

Figure 5-11 A), to the ultimate stress (axial plastic strain of 4%, Figure 5-11 B)

and on the onset of failure (axial plastic strain of 9%, Figure 5-11 C). It is worth

noting that the stress field is still homogenous for an axial plastic strain of 5%,

while at the onset of fracture (10% axial plastic strain), it is predicted that stress

localization and necking has occurred.

Figure 5-11. The von Mises stress distribution in the forged FEA specimen A) initial yielding, B) ultimate stress, and C) fracture.

Page 165: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

165

The predictive capabilities of the model have been assessed through

comparison with experimental results under uniaxial compression and uniaxial

tension with flat and axisymmetric specimens. In the following section, numerical

simulations of more complex loadings will be performed.

5.3.4 Predictive Capabilities of the New Distortional Hardening Model: Finite Element Analysis using Cyclic Loadings In order to illustrate the capability of the new hardening model for cyclic

loading, cyclic tension-compression simulations have been performed. The

specimen geometry is axisymmetric of diameter of 12.7 mm and height of 6.35

mm. Only one quarter of the specimen has been meshed with 3375 hexahedral

elements. A cyclic compression-tension loading controlled in displacement has

been applied to the specimen. The prescribed loading consists of an initial

compression 0.1u mm of the specimen follow by four cycles of tension-

compression loading 0.1u mm .

The predicted mechanical behavior under cyclic loading of the forged ES-

1 material is plotted in Figure 5-12. The absolute axial stress is plotted as a

function of equivalent plastic strain. It is important to notice the capability of the

new hardening model to capture the reduction in axial stress upon each load

reversal. This is due to the fact that the new hardening model depends on the

third invariant of plastic strain, 3J (see Chapter 4). For uniaxial tension loading,

3 0J , while for uniaxial compression loading, 3 0J .

Page 166: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

166

Figure 5-12. Evolution of the axial stress vs. equivalent plastic strain for tension-compression cyclic loading.

5.3.5 Predictive Capabilities: Finite Element Analysis of Free End Torsion

As discussed previously in Section 1.4, during monotonic free end torsion

loading of thin-walled cylindrical specimens, the occurrence of axial plastic strain

has been observed. It was demonstrated that the occurrence of these axial

plastic strains are due to the tension-compression asymmetry displayed by the

mechanical behavior of the material. In Chapter 4, it was deduced from the yield

criterion that the increment of plastic deformation (in the principal axis) under free

end torsion is:

Page 167: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

167

1

3222

23 322 3

30 0

2 33

32 0 03 2 3

230 0

3

P

c

JJJ c

F d cd d d

cJ cJ

ε (5-19)

According to Equation (5-19), the increment of thickness strain that

develops during free end torsion is proportional to 2

3

c. By symmetry and plastic

incompressibility, the axial strain is, therefore, directly related to the asymmetry

parameter, c. If c is zero (no tension-compression asymmetry), then pure shear

deformation occurs (i.e. no axial deformation). If the material has higher yielding

strength in uniaxial tension than in uniaxial compression (c>0), the specimen

walls thicken (positive thickness strain) and the length shortens (negative axial

plastic strain). The reverse holds true for greater yield stress in uniaxial

compression (c<0). For the ES-1 materials, irrespective of the processing, the

mechanical material has greater flow stress in uniaxial compression than in

uniaxial tension. Thus, it is expected that the specimen should elongate under

free end torsion.

Numerical predictions for free end torsion loadings have been performed.

The geometry of the specimen is shown in Figure 5-13 and was meshed with

4522 hexahedral elements (see Figure 5-14). The nodes at the lower extremity

of the specimen (y=0) were pinned, i.e., no displacement was allowed, while the

upper nodes (y=58.4) were tied to a rigid tool to impose torsion while ensuring

that all the upper nodes experienced the same boundary conditions. An angular

Page 168: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

168

rotation of 0.3 radian was applied to the rigid tool, but the axial displacement was

not constrained.

Figure 5-13. Geometry of the thin-walled specimen used for free-end torsion loading with dimensions expressed in millimeters.

Page 169: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

169

Figure 5-14. Isocontour of the predicted axial displacement that develops during free-end torsion loading of a forged ES-1 material:

A) undeformed and B) with 0.05 radians of axial rotation and axial displacement.

Figure 5-14 shows the axial displacement field for a rotation of the rigid

tool of 0.05 radian. For this angular displacement, an axial displacement of

0.047 mm is predicted. Therefore the specimen lengthened when subject to pure

rotation, a phenomenon known as the Swift effect. In Figure 5-15 is shown the

evolution of the axial strain with respect to the shear strain. The axial strain and

shear strain are calculated using

0

ln 1axial

u

L

and

0

γr

L

where r is the current radius, L0 is the initial length of the specimen, u is the axial

displacement, and is the twist angle. Due to the fact that the yield stress in

uniaxial compression is larger than the one in uniaxial tension, it is predict that

the ES-1 materials will elongate.

Page 170: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

170

Figure 5-15. Longitudinal elongation by shear strain for forged quasi-static specimens under quasi-static torsion.

5.4 Comparison between the Model and Dynamic Characterization Data

In this section, the model developed previously will be applied to high

strain-rate loading and impact testing. Using the dynamic implicit solver of

ABAQUS, the model predictions were compared with experimental data obtained

with a SHPB apparatus and Taylor impact tests.

5.4.1 Comparison between Model and Dynamic Characterization Data

To characterize the dynamic behavior of the ES-1 materials, SHPB

compressive experiments were performed and discussed in Section 3.3. In this

section, the numerical prediction will be compare with the experimental results.

But, in this section, we will not compare the strain-stress curve obtained at high-

strain rate, but the signals that were recorded by strain-gages located on the

incident and transmitted bars. Therefore, the SHPB system was modeled in its

entirety (i.e. the striker bar, the incident and transmitted bars as well as the

specimen) and wave propagations through the bars were studied. The geometry

of each component of the UF REEF SHPB system is presented in Chapter 3.

Page 171: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

171

Due to symmetry conditions, only one quarter of the bars were meshed. The

total size of the mesh is 6942 elements. Close to the location of the strain-gages

on the incident bar and the transmitted bars, the mesh is refined to obtain a fine

wave propagation signal. Contacts are defined between each of the system

components, and the specimen is initially in contact with the incident and

transmitted bars (see Figure 5-16). The initial velocity of the striker bar is 14 m/s.

As the striker bar impacts the incident bar, an elastic wave is created and further

propagates through the SHPB system and the cylindrical specimen of ES-1. It is

worth noting that the bars are considered to only deform elastically, while the

specimen is described by the developed elastic plastic model.

Figure 5-16. The finite element SHPB and an inset containing a close-up of an ES-1 specimen.

The predicted strain waves, i.e. the incident, reflected and transmitted

waves, are compared with the experimental ones (see Figure 5-17). Note the

very good agreement between the experimental data and the predictions. The

Page 172: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

172

evolution of the strain wave with time is correctly predicted by the model. The

measured strain waves from the finite element simulation and the experiment

were compared in Figure 5-17 to validate the rate dependent modeling of Eglin

steel.

Figure 5-17. SHPB incident, transmitted, and reflected waves from forged specimen and FEA.

5.4.2 Model validation: Simulation of Taylor Impact Tests

To further validate the prediction of the developed model for high strain-

rate, finite element simulations of Taylor impacts have been performed. The

Taylor impact test developed by Taylor (1948) consists of launching a solid

cylindrical specimen at an elevated velocity of the order of 100 to 300 m/s

against a stationary rigid anvil. When the specimen impacts the rigid stationary

anvil an elastic compressive wave is generated at the impact interface and

travels back and forth through the specimen. For a sufficiently high impact

velocity, for which the magnitude of the compressive wave reaches the yield

Page 173: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

173

stress of the material, the impact end undergoes plastic deformation. Therefore,

the plastic front starts propagating from the impact interface. However, only a

portion of the specimen deforms plastically.

Torres et al. (2009) completed ten Taylor impact tests of cylindrical ES-1

specimens. Torres et al. used the analyses of the deformed specimens

developed by Jones et al. (1998). The initial diameter of the specimens was d0

=4.17 mm while its initial length was L0 = 31.317 mm. The impact velocities

varied between 148 and 200 m/s. As already mentioned, when recovered, the

length of the deformed Taylor impact specimens, Lf, is less than than its initial

length, L0. Furthermore, the recovered specimen consists of a rear portion of

length (lf) with a constant cross sectional area (A0), i.e. that has only deformed

elastically, and a portion that has deformed plastically in which the cross-section

(A) gradually increases as shown in Figure 5-19. The analysis depends entirely

upon three measurements of the deformed specimens: the cross sectional area

at several locations along the length of the specimen, the length from the rear of

the projectile to the corresponding cross sectional area (lf), and the total

deformed length of the projectile (Lf). Using plastic incompressibility, the

engineering plastic strain was calculated based on the variation of the cross-

sectional area.

0 1A

eA

(5-20)

As shown in Figure 5-19, Torres et al. (2009) reported the location of the

plane (lf/L0) of the cross section of the recovered specimen that was subjected to

engineering plastic strain e={2, 3, 4, 5, ..12%}. For more details about the

Page 174: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

174

analysis of the Taylor impact experiments, the reader is referred to Jones et al.

(1998).

FEA simulations using the developed model in conjunction with the implicit

solver of ABAQUS have been performed for the ES-1 material at an impact

velocity of 148 m/s. Due to the symmetry of the problem, only a quarter of the

specimen was meshed with 38880 hexahedral elements. The mesh is refined in

the impact zone to accurately capture the deformation of the impacted surface of

the specimen. Boundary conditions were applied such as to maintain the

symmetry of the problem. The simulations were performed in two steps,

associated with the specimen launch and impact, respectively. In the first step,

the impact velocity was applied to all of the nodes such as to represent the free

flight or launch of the specimen. In the second step, the impact of the specimen

with the rigid anvil was reproduced by imposing a null forward velocity only to the

nodes belonging to the impact surface while the other nodes were not

constrained anymore. In this manner, the test conditions were reproduced with

fidelity. Furthermore, it is assumed that the impact ends when the axial velocity

at the end of the specimen becomes null. Figure 5-18 presents the equivalent

plastic strain isocontour in the Taylor impact specimen of forged ES-1.

Page 175: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

175

Figure 5-18. Equivalent plastic strain isocontour in the cylindrical Taylor impact specimen of forged ES-1.

The FEA predictions are in good agreement with the experimental data.

For an impact velocity of 148 m/s, the predicted normalized final length of the

specimen is 0

0.954fL

L

, while the reported experimental data by Torres et al.

(2009) is 0

0.957fL

L

. To directly compare experimental data and FEA

predictions, the axial location of the cross-section subjected to a plastic strain of

2.54,3.28,10.35%e have been reported in the plane (lf/L0, Lf/L0) in Figure 5-19.

It is worth noting the good agreement between the model prediction and

experiments.

Page 176: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

176

Figure 5-19. Linear relationship of deformed section length by deformed total length following experimental and simulated Taylor impact tests of specimens of ES-1.

Using the one dimensional wave analysis of Jones et al. (1998) of the

deformed specimen dimensions, Torres et al. presented the stress-strain

response shown in Figure 5-20 for strain rates between 103 and 104 s-1. It is also

worth comparing the experimental results with the numerical predictions.

Page 177: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

177

Figure 5-20. Stress strain response of ES-1 subject to dynamic strain rates in Taylor impact tests conducted by Torres et al.

The excellent agreement between the simulated and experimental data for

the Taylor impact tests assesses the predictive capabilities of the developed

model of ultra high strength martensitic steels under dynamic loadings.

In this chapter, FEA predictions of the developed model have been

performed for monotonic tests (e.g. uniaxial tension and free-end torsion), cyclic

tension-compression loadings, and dynamic events (SHPB and Taylor impact

tests). A good overall agreement between the prediction and available

experimental data was observed.

Chapter 6 includes a summary of this dissertation and research as well as

recommendations for future research.

Page 178: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

178

CHAPTER 6 CONCLUSIONS AND RECOMMENDATIONS

In this dissertation an integrated approach using experimental, theoretical,

and numerical techniques was employed to further the understanding of ultra

high strength martensitic steels. The overall goal was to determine, for the first

time, the effect on the mechanical response of ES-1 steel of three different

manufacturing processes, namely forging, casting, and HIPing. Prior to the

experimental characterization, the ES-1 materials were subject to microscopic

analysis. From the SEM observations of the ES-1 material texture, it was

concluded that processing did not introduce anisotropic features. Irrespective of

the processing, the observed texture was random indicating that the materials

are isotropic.

To characterize the mechanical behavior, a suite of mechanical tests were

conducted. The quasi-static tests 310 performed were monotonic

compression, monotonic tension, and cyclic tension tests.

As concerns the monotonic tests, first axisymmetric specimens (circular

cross section) with a 25.4 mm gage length were tested in uniaxial tension. The

average yield strength for the forged, cast and HIP’d, and cast axisymmetric

specimens using 0.2% offset method were 1541 MPa, 1414 MPa, and 1445

MPa, respectively. The strain-to-failure was of 10%, 8%, and 13% for the forged,

cast, and cast and HIP’d materials, respectively. Irrespective of processing, the

posttest cross section was circular. It was thus confirmed by mechanical testing

that the ES-1 materials studied are isotropic.

Page 179: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

179

Moreover, for monotonic tensile loading the influence of the specimen

geometry on localization of the deformation and strain to failure was investigated.

Both global and local strain measurements using DIC were done. For DIC, two

orthogonal cameras were used such as to allow measurements of both the width

and thickness strains. Irrespective of the processing history of the material,

these strains were equal thus confirming the isotropy of ES-1. The true stress-

true strain behavior prior to strain localization was found to be very close to that

of the round axisymmetric specimens, but the strain-to-failure of the flat

specimens was about half that of the axisymmetric specimens.

Quasi-static uniaxial compression tests were also conducted on cylindrical

specimens with a length and diameter of 7.62 mm. It was found that ES-1

exhibits similar yield values in tension and compression. As the plastic strain

accumulates, however, the plastic flow stress in compression exceeds that in

tension. At 4% axial strain, the ratio of flow stress in tension to compression

approaches 0.89.

To elucidate the evolution of the asymmetry of the yield surface, cyclic

tests on ES-1 axisymmetric specimens with a 25.4 mm gage length were

conducted using an MTS load frame at the UF REEF. Under tensile strain, the

tensile flow stress increased while the compressive flow stress was diminished.

Under compressive strain the opposite trend was observed. It was concluded

that the evolution of asymmetry of the yield surface is a function of the plastic

strain tensor.

Page 180: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

180

As concerns the dynamic tests, cylindrical specimens with a length and

diameter of 5.08 mm were tested in a SHPB system with a 19 mm diameter at

the UF REEF. A gas gun was used to accelerate a striker bar to impact the

incident bar at approximately 13 m/s creating a strain wave that deformed the

specimen at an average strain rate of 700 s-1. Irrespective of the processing

history, the hardening behavior of ES-1 subjected to high-rate strain was similar

to that under quasi-static loadings, but the strength was higher by a dynamic

increase factor of 1.1.

To describe the experimentally observed mechanical behavior a new

model was developed in the framework of the theory of viscoplasticity. To

account for tension-compression asymmetry in yielding the isotropic form of

Cazacu and Barlat (2004) yield criterion was used in conjunction with a new

hardening model. The key novel aspect of this new hardening model is that it

accounts for the distortion of the yield locus for both monotonic and cyclic

loadings. The noteworthy aspect is that the model is dependent on both

invariants of the plastic strain, and it is isotropic in nature. This is a departure

from the current practice where in order to account for Bauschinger effects an

additional tensorial internal variable, namely the overstress, is introduced. The

prime advantage of the proposed hardening model is that it is accurate under

monotonic loading, but also describes the Bauschinger effect.

Because of the slightly increased accuracy under monotonic loading and

significantly improved characterization of hardening under changes in strain path,

it is believed that the new model will provide more accurate results under

Page 181: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

181

complex loadings. For this purpose, the capabilities of the new hardening model

were illustrated for combined loadings such as tension-shear; compression-

shear. The tests necessary for the determination of the parameters involved in

the new hardening model and the identification procedure were outlined.

Specifically, the dependence of hardening on the second invariant of strain can

be isolated and thus fully identified from pure shear tests. Such a law induces a

growth of the yield surface in the deviatoric plane. The strain-hardening

dependence on the third invariant of plastic strain cannot be isolated.

Nevertheless, from uniaxial tensile and compressive test data when strain-

hardening depends on both invariants, assuming linear decomposition, it

becomes possible to identify the dependence on the second invariant of the

plastic strain. Since the third invariant of plastic strain is not associated with

growth of the entire yield surface, to accommodate the prescribed deformation in

two-stage tests involving uniaxial load reversal, in the octahedral plane, the yield

surface normal at the six states associated to pure shear strain ought to rotate.

Therefore, it is predicted that in order to accommodate Bauschinger effects the

yield surface is distorted. For example, under uniaxial tensile strain the initially

“triangular” yield surface will extend along the tensile meridians and will contract

along the compressive ones (i.e. the rounded corners of the rounded “triangle”

intersect the tensile meridians). Under uniaxial compressive strain, the opposite

distortion occurs. Thus, the yield surface evolves differently depending on the

strain history (tension followed by compression, or compression followed by

tension).

Page 182: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

182

This new model was implemented in a fully three dimensional, implicit

finite element code and then validated using independent experimental data.

Specifically, for quasi-static loading validation was done for pure shear loadings

while for dynamic loadings the Taylor impact loadings were simulated.

For comparison purposes, the von Mises yield criterion in conjunction with

a distortional model involving only dependence on the third invariant of the plastic

strain, and the Cazacu and Barlat (2004) yield criterion with evolving hardening

depending only on the accumulated plastic strain was also applied to ES-1

materials. The same experimental data was used for identification of these

formulations. Both formulations were also implemented in the same FE solver

(ABAQUS) using a UMAT. FE simulations of flat ES-1 specimens under uniaxial

tensile loading were conducted and a side-by-side comparison of the strain maps

using DIC in the experimental characterization and both FE analyses was

presented. It was shown that the Cazacu and Barlat (2004) criterion coupled with

the full distortional hardening model provides the most accurate representation of

the test results.

In addition, verification of the FE implementation and of the capabilities of

the model was done by simulating free-end torsion of a thin-walled cylinder. Due

to the tension-compression asymmetry of the material, elongation under torsion

was predicted.

Most importantly, this dissertation research provides new insights and

understanding of plastic deformation under combined loadings. While the new

model developed was applied to ES-1 materials, it has a much broader range of

Page 183: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

183

validity. Future research involving the use of this hardening model in conjunction

with anisotropic yield criteria and its applications to strongly textured materials

should be of great interest.

Page 184: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

184

LIST OF REFERENCES

1. ABAQUS, 2009. User’s Manual for Version 6.8, vol. I-V. Dassault Systemes

Simulia Corp., Providence, RI.

2. ASTM Standard E111 -04, “Standard Test Method for Young’s Modulus,

Tangent Modulus, and Chord Modulus,” ASTM International, West

Conshohocken, PA 2015, DOI: 10.1520/E0111-04R10, www.astm.org.

3. ASTM Standard E112-13, “Standard Test Methods for Determining Grain

Size,” ASTM International, West Conshohocken, PA 2015, DOI:

10.1520/E112-13, www.astm.org.

4. ASTM Standard E8/E8M-15a, “Standard Test Methods for Tension Testing

of Metallic Materials,” ASTM International, West Conshohocken, PA 2015,

DOI: 10.1520/E0008_E0008M-15A, www.astm.org.

5. ASTM Standard E9 -09, “Standard Test Methods for Compression Testing

of Metallic Materials at Room Temperature,” ASTM International, West

Conshohocken, PA 2015, DOI: 10.1520/E0009-09, www.astm.org.

6. ASTM Standard E9 -09, “Standard Test Methods for Compression Testing

of Metallic Materials at Room Temperature,” ASTM International, West

Conshohocken, PA 2015, DOI: 10.1520/E0009-09, www.astm.org.

7. Atkinson, H. V., Davies, S., 2000. Fundamental aspects of hot isostatic

pressing: an overview. Metall. Mater. Trans. 31A, 2981-3000.

8. Bau, Y., Wierzbicki, T., 2004. On fracture locus in the equivalent strain and stress triaxiality space. International Journal of Mechanical Sciences 46, 81-98.

9. Bauschinger, J., 1886. Mitt. mech. tech. lab. Munchen, Heft 13,1.

10. Belytschko, T., Liu, W. K., Moran, B., & Elkhodary, K., 2013. Nonlinear finite

elements for continua and structures. John Wiley & Sons, New York.

11. Bhadeshia, H., Honeycombe, R., 2006. Steels – microstructure and

properties, 3rd ed. Elsevier. 12. Billington, E.W., 1976. Nonlinear mechanical response of various metals: II.

Permanent length changes in twisted tubes. J. Phys. 9, 533-552.

Page 185: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

185

13. Blewitt, T.H., Coltman, R.R., Redman, J.K., 1957. Low-temperature

deformation of copper single crystals. J. Appl. Phys. 28, 651. 14. Boyce, B.L., Dilmore, M.F., 2009. The dynamic tensile behavior of tough,

ultra high-strength steels at strain rates from .0002 s-1 to 200 s-1. International Journal of Impact Engineering 36, 263-271.

15. Bridgman, P.W., 1952. Studies in large plastic flow and fracture: with

special emphasis on the effects of hydrostatic pressure. McGraw-Hill, New York, NY.

16. Brownrigg, W.A., Spitzig, W.A., Richmond, O., Teirlinch, D., Embury, J.D.,

1983. The influence of hydrostatic pressure on the flow stress and ductility

of a spherodized 1045 steel. Acta Metall. 31, 1141-1150.

17. Cazacu, O., Barlat F., 2004. A criterion for description of anisotropy and

yield differential effects in pressure-insensitive metals. Int. J. Plas. 20, 2027-

2045.

18. Cazacu, O., Plunkett, B., Barlat, F., 2006. Orthotropic yield criterion for

hexagonal closed packed metals, Int. J. Plas. 22, 1171-1194.

19. Cazacu, O., Revil-Baudard, B., Barlat, F., 2012. New interpretation of

monotonic Swift effects: role of tension-compression asymmetry,

Mechanics of Materials 57, 42-52.

20. Cazacu, O, Stewart, J.B., 2009. Plastic potentials for porous aggregates with matrix exhibiting tension-compression asymmetry. J. Mech. Phys. Solids 57, 325-341.

21. Chait, R., 1971. Factors influencing the strength differential of high strength

steels. Defense Technical Information Center, AMMRC TR 71-48.

22. Chree, C., 1889. The equation of anisotropic elastic solid in polar and

cylindrical coordinates, their solution and applications. Transactions of

Cambridge Philosophical Society 14,250-369.

23. Collins, J.A., 1993. Failure of materials in mechanical design: analysis,

prediction, prevention. John Wiley & Sons, New York, NY.

24. Cristescu N., 1967. Dynamic plasticity. North-Holland Publishing Company-

Amsterdam. John Wiley & Sons, New York.

Page 186: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

186

25. Dafalias, Y.F., Popov, E.P., 1975. Plastic internal variables formalism of

cyclic plasticity. J. App. Mech. 43, 645.

26. Davies R.M., 1948. A critical study of the Hopkinson pressure bar, Philos.

Trans. R. Soc. 240, 375-457.

27. Davies E.D., Hunter S.C., 1963. The dynamic compression testing of solids

by the method of the split Hopkinson pressure bar. J. Mech. Phys. Solids

11, 155-179.

28. Demeri, Mahmoud Y., 2013. Advanced high-strength steels-science,

technology, and application. ASM International.

29. Dilmore, M., Ruhlman, J. D., 2009. Eglin steel – a low alloy high strength

composition. USPTO US7537727.

30. Drucker, D.C., 1979. Relation of experiments to mathematical theories of

plasticity, J. Appl. Mech. 16, 349-357.

31. Faleskog, J., and Barsoum, I., 2013. Tension-torsion fracture experiments –

Part I: Experiments and a procedure to evaluate the equivalent plastic

strain. International Journal of Solids and Structures 50, 4241-4257.

32. Fedchun, V., Razdolsky, M., Dilmore, M., 2003. Steel with high strength and

toughness. Advanced Materials and Processes, 33-36.

33. Foley, R. D., Griffin, J.A., and Monroe, C. A., 2016. Characterization of a

cast low alloy, ultra high strength martensitic steel. Proceedings of the 70th

SFSA Technical and Operating Conference. Steel Founders Society of

America, Chicago, IL.

34. Follansbee, P. S., Frantz, C., 1983. Wave propagation in the split

Hopkinson pressure bar. Journal of Engineering Materials and Technology

105, 61-66.

35. Frew, D. J., Forrestal, M. J., Chen, W., 2005. Pulse shaping techniques for

testing elastic-plastic materials with a split Hopkinson pressure bar. Society

for Experimental Mechanics 45, 186-195.

36. Goldsmith, W., 1960. Impact: the theory and physical behavior of colliding

solids, Edward Arnold Ltd. London, UK.

Page 187: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

187

37. Gong, J. C., Malvern, L. E., Jenkins, D. A., 1990. Dispersion investigation in

the split Hopkinson pressure bar. Journal of Engineering Materials and

Technology 112, 309-314.

38. Gray, G.T., 2000. Classic split-Hopkinson pressure bar testing. ASM

Handbook Volume 8, Mechanical Testing and Evaluation (ASM

International), 462-476.

39. Green, A. E. and Naghdi, P.M., 1965. A general theory of an elastic-plastic

continuum. Archive Rational Mech. & Anal. 18, 251-281.

40. Hershey, A.V., 1954. The plasticity of an isotropic aggregate of anisotropic

face centered cubic crystals. J. Appl. Mech. Trans. ASME 21, 241-249.

41. Hill, R., 1987. Constitutive dual potentials in viscoplasticity. J. Mech. Phys.

Solids 35, 35-33.

42. Hill, R., 1948. A theory of the yielding and plastic flow of anisotropic metals.

Proc. R. Soc. Lond. 193, 281-297.

43. Hirth, J.P., Cohen, M., 1970. On the strength-differential phenomenon in

hardened steel. Metall. Trans. 1, 3-8.

44. Hodge, P.G., Jr. 1957. A general theory of piecewise linear plasticity based

on maximum shear. J. Mech. Phys. Solids 5, 242-260.

45. Hopkinson B., 1914. A method of measuring the pressure produced in the

detonation of high explosives or by the impact of bullets, Philos. Trans. R.

Soc. A 213, 437-456.

46. Hopson, M.V., Scott, C.M., Lambert, D., 2015. Applicability of statistical flaw

distributions of Eglin steel for fracture calculations. Procedia Engineering

103, 213-220.

47. Hosford, W.F., 1972. A generalization isotropic yield criterion. J. Appl. Mech.

39, 607-609.

48. Hosford W.F., 1993. The mechanics of crystals and textured polycrystals.

Oxford University Press, New York.

Page 188: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

188

49. House, J.W., 1989. Taylor impact testing. Technical Report AFATL-TR-89-

41, AD-A215 018.

50. Huang, J.Y., Wu, Y.K., Ye, H.Q., 1996. Deformation structures in ball milled

copper. Acta Metall. 44, 1211-1221.

51. Hughes, T., 1984. Numerical implementation of constitutive models: rate

independent deviatoric plasticity. in: S. Nemat-Nasser et al. (Eds.),

Theoretical foundations for large-scale computations for nonlinear material

behavior. Martinus Nijhoff, Dordrecht, 29–57.

52. Johnson, G., Cook, W., 1983. A constitutive model and data for metals

subjected to large strains, high strain rates, and high temperatures.

53. Jones, S.E., Drinkard, J.A., Rule, W.K., Wilson, L.L., 1998. An elementary

theory for the Taylor impact test. International Journal of Impact Engineering

21, 1-13.

54. Jones, S.E., Maudlin, P.J., Foster, F.C., 1996. An engineering analysis of

plastic wave propagation in the Taylor test, International Journal of Impact

Engineering 19, 95-106.

55. Kahloun C., Le L., Franciosi P., Chavane M.H., Ait, E., Monnet, G., 2012.

Topological analysis of (110) slip in alpha-iron crystal from in-situ tests in

atomic force microscope. Acta Mater. 61, 6453-6455.

56. Kocks U.F., 2000. Anisotropy and symmetry, in: Kocks, U.F., Tome, C.N.,

Wenk, H.R., Texture and anisotropy. Cambridge University Press,

Cambridge, UK, 30.

57. Kolsky H., 1949. An investigation of the mechanical properties of materials

at very high rates of loading. Proc. Phys. Soc. 62B, 676-700.

58. Krieg, R.D., 1975. A practical two-surface plasticity theory. J. App. Mech.

47, 641.

59. Lambert, D.E., Weiderhold, J., Hopson, M.V., Osborn, J., 2010. Controlled

loading fragmentation: experiments and continuum damage modeling.

Defense Technical Information Center.

Page 189: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

189

60. Landgraf, R.W., 1970. The resistance of metals to cyclic deformation.

Achievement of high fatigue resistance in metals and allows. ASTM STP

467, 3-36.

61. Last, H.R., Garrett, R.K., and Rajendran, A.M., 1996. A comparative study

of high strain rate behavior of three martensitic steels. American Institute of

Physics 631-634.

62. Leister, B.M., DuPont, J.N., Watanabe, M., Abrahams, R.A., 2015.

Mechanical properties and microstructural evolution of simulated heat-

affected zones in wrought Eglin steel, Metall. Mater. Trans. A 46A, 5727-

5746.

63. Leslie, W.C., Sober, R.J., 1967. The strength of ferrite and of martensite as

functions of composition, temperature, and strain rate. ASM Trans. 60, 459.

64. Little, C.D., Machmeier, P.M., 1979. High strength fracture resistant

weldable steels. USPTO 4076525.

65. Lubliner, J., 1990. Plasticity theory. Dover Publications Inc., Mineola, NY.

66. Lynch, P.C., 2011. The development of ultrahigh strength low alloy cast

steels with increased toughness. ProQuest LLC., Ann Arbor, MI.

67. Maiden, C.J., Green, S.J., 1966. Compressive strain-rate tests on six

selected materials at strain rates from 10-3 to 104 in/in/sec. J. Appl. Mech.

33, 496-504.

68. Malvern, L., Jenkins, D. A., Jerome, E., Gong, J. C., 1987. Disperson

correction for split-Hopkinson pressure bar data. Defense Technical

Information Center.

69. Malvern, L., 1969. Introduction to the mechanics of a continuous medium.

Prentice-Hall Inc., Upper Saddle River, NJ.

70. Martin, B.E., Flater, P.J., Abrahams, R., Neel, C., Reinhart, W., Chhabildas,

L., 2012. Dynamic characterization of Eglin steel by symmetric impact

experimentation. AIP Conference Proceedings 1426, 979.

Page 190: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

190

71. Maudlin, P.J., Gray, T.G., Cady, C.M., Kaschner, C.G., 1999. High rate

material modelling and validation using the Taylor cylinder impact test.

Philos. Trans. R. Soc. 357, 1707-1729.

72. Melan, E., 1938. Zur plastizitat des raumlichen kontinuums. Arch. Appl.

Mech. 9, 116-126.

73. Mishra A., Kad B., Gregori F., Meyers M., 2007. Microstructural evolution in

copper subjected to severe plastic deformation: experiments and analysis.

Acta Materialia 55, 13-28.

74. Nadai, A., 1963. Theory of flow and fracture of solids. McGraw-Hill, New

York, NY.

75. Nagdi, P., 1960. Stress-strain relations in plasticity and thermoplasticity. In:

Proceedings of the 2nd symposium on naval structural mechanics.

Pergamon, London, UK.

76. Paul, B., 1968. Macroscopic criteria for plastic flow and brittle fracture. In

Fracture-a treatise, Liebowitz, Academic Press, London, UK, 315-496.

77. Pochhammer, L., 1876. On the propagation velocities of small vibrations in

an infinite isotropic cylinder. Zeit Schrift fur Reine und Ange wandte

Mathematik 81, 324-336.

78. Prager, W., 1956. A new method for analyzing stress and strain in work

hardening, plastic solids. J. Appl. Mech. 23, 493.

79. Revil-Baudard, B., Chandola, N., Cazacu, C., and Barlat, F., 2014.

Correlation between swift effects and tension compression asymmetry in

various polycrystalline materials. J. Mech. Phys. Solids 70, 104-115.

80. Rice, J.R., 1976. The localization of plastic deformation, Theoretical and

Applied Mechanics (Proceedings of the 14th International Congress on

Theoretical and Applied Mechanics, Delft, 1976, ed. W.T. Koiter) North

Holland Publishing Co. 207-220.

81. Rice, J.R., Rudnicki, J.W., 1979. A note on some features of the theory of

localization of deformation. International Journal of Solids and Structures

16, 597-605.

Page 191: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

191

82. Ross, A.C., 1986. Split-Hopkinson pressure bar tests. Defense Technical

Information Center, ESL-TR-88-82.

83. Sharpe, W.N., Hoge, K.G., 1972. Specimen strain measurements in the split

Hopkinson pressure bar experiment. Experimental Mechanics 12, 570-574.

84. Simo, J. C., Hughes, T. J. R., 1998. Computational inelasticity. Springer-

Verlag, New York, NY.

85. Smith, R.L., Sandland, G.E., 1922. An accurate method of determining the

hardness of metals, with particular reference to those of a high degree of

hardness. Proc. Inst. Mech. Eng. 1, 623-641.

86. Spitzig, W.A., Sober, R.J., Richmond, O., 1976. The effect of hydrostatic

pressure on the deformation behavior of maraging and HY-80 steels and its

implications for plasticity theory. Metall. Trans. A, 1703-1710.

87. Spitzig, W.A., Sober, R.J., Richmond, O., 1975. Pressure dependence of

yielding and associated volume expansion in tempered martensite. Acta

Metall. 23, 885-893.

88. Stewart, J., 2009. Development of yield criteria for describing the behavior

of porous metals with tension-compression asymmetry. ProQuest LLC., Ann

Arbor, MI.

89. Susan, D. F., Crenshaw, T.B., and Gearhart, J.S., 2015. The effects of

casting porosity on the tensile behavior of investment cast 17-4PH stainless

steel, Journal of Materials Engineering Performance 28, 2917-2924.

90. Sutton M., Orteu J., Schrier H., 2009. Image correlation for shape, motion,

and deformation measurements: basic concepts, theory, and applications.

Springer Publishing, New York, 120-137.

91. Swift, H., 1947. Length changes in metals under torsional overstrain.

Engineering 163, 253-257.

92. Swift, H.W., 1952. Plastic instability under plane stress. J. Mech. Solids 1,1-

18.

93. Takeuchi T., 1969. Temperature dependence of work-hardening rate in iron

single crystals. Journal of the Physical Society of Japan 26, 354-362.

Page 192: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

192

94. Takeuchi T., 1969b. Orientation dependence of work-hardening in iron

single crystals. Japanese Journal of Applied Physics 8, 320-328.

95. Taylor, G., 1948. The use of flat-ended projectiles for determining dynamic

yield stress. R. Soc. Lond. 194, 289-299.

96. Torres, K. L., Clements, H. A., Jones, S. E., Dilmore, M., Martin, B., 2009.

Dynamic strength estimates for a high-strength experimental steel. Journal

of Pressure Vessel Technology 131, 1-6.

97. Van Aken, D.C., Webb, T.O., Green, J. D., Grain refinement strategies in

high strength cast steel.

98. Voce, E., 1948. The relationship between stress and strain for

homogeneous deformation. J. Inst. Metals 74, 537-562.

99. Wang G., Strachan A., Cagin T., Goddard III W., 2004. Calculating the

Peierls energy and Peierls stress from atomistic simulations of screw

dislocation dynamics: application to bcc tantalum. Model. Simul Mater. SC,

12, S371–S38.

100. Wang B., Zhang J., Lu G., 2003. Taylor impact test for ductile porous

materials part I: experiments. International Journal of Impact Engineering

28, 499-511.

101. Webb, T. O., Van Aken, D. C., Lekakh, S. N., 2014. Evaluating chemical

homogeneity in the performance of Eglin steel. American Foundry Society

Proceedings 14-017.

102. Weiderhold, J, Lambert, D.E., Hopson, M., 2010. Experimental design and

data collection for dynamic fragmentation experiments. Proceedings of

ASME Pressure Vessels and Piping Division.

103. Yadav, S., Chichili, D.R., Ramesh, K.T., 1995. The mechanical response of

a 6061-T6 Al/Al203 metal matrix composite at high rates of deformation.

Acta Metall. 43, 4453-4464.

104. Yeganeh, M., Naghdabadi, R., 2006. Axial effects investigation in fixed-end

circular bars under torsion with a finite deformation model based on

logarithmic strain. International Journal of Mechanical Sciences 48, 75-84.

Page 193: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

193

105. Zerilli, F.J., Armstrong, R.W., 1988. Dislocation mechanics based

constitutive relations for material dynamics modeling: slip and deformation

twinning in iron. In: Shock Waves in Condensed Matter. In: Smidth, S.C.,

Holmes, N.C. (Eds.). Elsevier, Amsterdam.

Page 194: EXPERIMENTAL CHARACTERIZATION AND MODELING OF HIGH

194

BIOGRAPHICAL SKETCH In May of 2000, Elizabeth Kay Bartlett graduated with a Bachelor of

Science in mechanical engineering from Mississippi State University. As a

graduate research assistant to Dr. Susan Hudson, she continued her education

and completed a Master of Science in mechanical engineering in May of 2002.

While working as a weapons test engineer for the United States Air Force

(USAF), Elizabeth became interested in the fragmentation of warheads in arena

tests. In May 2005, her only daughter, Gwyneth Webb, was born at Fort Walton

Beach Medical Center. While working full-time, she began taking classes at the

UF Research and Engineering Education Facility in Shalimar, FL. In August of

2016, she was awarded the Science, Mathematics, and Research for

Transformation (SMART) scholarship to complete her Doctor of Philosophy in

mechanical engineering under the advisement of Prof. Oana Cazacu.