Sie sind auf Seite 1von 336

Solute Enhanced Strain Hardening of Aluminum Alloys

to Achieve Improved Combinations of Strength and Toughness



A Thesis

Submitted to the Faculty

of

Drexel University

by

Christopher James Hovanec

in partial fulfillment of the

requirements for the degree

of

Doctor of Philosophy

April, 2011


























Copyright 2011
Christopher J. Hovanec. All Rights Reserved.


ii
Dedications
This body of work is dedicated to the memory of my maternal grandfather,
Michael Gerald Conway (April 27, 1922 - June 12, 2005). My grandfather always
stressed the importance of working hard, getting an education, and doing what is right.
These lessons have served me well.

iii
Acknowledgments
Foremost, I would like to thank my advisors, Prof. Roger D. Doherty and Prof.
Surya R. Kalidindi. The guidance and support provided by these gentlemen has been
instrumental in the successful completion of this project as well as in my professional
development.
I would like to express my gratitude to Mark Shiber and Joseph Bilko for their
assistance in fabricating the specialty tooling, fixturing, and specimen required to
complete this project. I would also like to thank Robert Taylor for his assistance in
configuring the experimental setups and data collection systems utilized in this project.
I am thankful to Prof. Richard Knight for the organizational and laboratory
management skills which he helped me to develop.
I am grateful to all the past and present members of the Mechanics of
Microstructures Group for their support and encouragement. I would like to specifically
acknowledge Dejan Stojakovic whom I have shared many hours in the laboratory with, of
which I have many found memories.
The alloys used throughout this project were provided by Alcoa. Quenching and
heat treating supplies were provided by Houghton International.
My graduate studies and much of this effort were funded by the Air Force Office
of Scientific Research (AFOSR) under the Metals Program (FA9550-04-1-0018).
The Naval Air Warfare Center - Aircraft Division at Naval Air Station Patuxent
River provided access to their mechanical testing equipment/facilities that allowed for the
nonlinear fracture toughness testing to be performed. This was made possible by the
efforts of Darrel Tenney, Dr. Eui Lee, and AIR-4.3.4.1.

iv
Table of Contents

List of Tables ................................................................................................................... viii

List of Figures..................................................................................................................... x

CHAPTER 1. INTRODUCTION....................................................................................... 1

CHAPTER 2. BACKGROUND......................................................................................... 8

2.1 Plastic Deformation................................................................................................... 8
2.2 Strengthening Mechanisms ..................................................................................... 15
2.2.1 Strain Hardening ............................................................................................... 16
2.2.2 Grain Boundary Strengthening (Hall-Petch Effect) .......................................... 21
2.2.3 Solution Hardening ........................................................................................... 29
2.2.4 Precipitation Hardening..................................................................................... 37
2.2.5 Total Strengthening Effect ................................................................................ 47
2.3 Material Failure by Fracture.................................................................................... 48
2.4 Fracture Toughness ................................................................................................. 51
2.4.1 Griffith Energy Balance and Energy Release Rate ........................................... 53
2.4.2 Stress Intensity Approach.................................................................................. 58
2.4.3 Plane Stress vs. Plane Strain ............................................................................. 62
2.4.4 Nonlinear Fracture Mechanics (J-Integral) ....................................................... 65
2.5 Relationship between Strength and Toughness....................................................... 68
2.6 Metallurgical Aspects of Fracture ........................................................................... 71
CHAPTER 3. MATERIAL............................................................................................... 80

3.1 Aluminum Alloys .................................................................................................... 80
3.1.1 AA2524............................................................................................................. 85

v
CHAPTER 4. EFFECT OF GP ZONE FORMATION ON STRAIN HARDENING
RATE................................................................................................................................ 88

4.1 Natural Aging .......................................................................................................... 91
4.1.1 Solution Heat Treating ...................................................................................... 92
4.1.2 Excess Vacancies .............................................................................................. 93
4.1.3 Guinier-Preston Zone Formation ...................................................................... 95
4.1.4 Residual Stress .................................................................................................. 97
4.2 Strain Path Change ................................................................................................ 101
4.3 Processing and Experimental Method................................................................... 104
4.3.1 Heat Treating................................................................................................... 104
4.3.2 Deformation Processing - Cold Rolling.......................................................... 107
4.3.3 Mechanical Testing ......................................................................................... 108
4.4 Experimental Results............................................................................................. 110
4.4.1 Heat Treating................................................................................................... 110
4.4.1.1 Quench Rate.............................................................................................. 110
4.4.1.2 Origin of Mid-Plane Cracking .................................................................. 116
4.4.1.3 Uphill Quenching...................................................................................... 117
4.4.2 Deformation Processing - Cold Rolling.......................................................... 117
4.4.2.1 Hardness and Conductivity....................................................................... 117
4.4.2.2 Strengthening and SHR Data.................................................................... 124
4.4.3 Compression Testing Data .............................................................................. 129
4.5 Discussion.............................................................................................................. 136
4.5.1 Heat Treating................................................................................................... 136
4.5.1.1 Quench Rate.............................................................................................. 136

vi
4.5.1.2 Origin of Mid-Plane Cracking .................................................................. 145
4.5.1.3 Uphill Quenching...................................................................................... 146
4.5.2 Deformation Processing - Cold Rolling.......................................................... 148
4.5.2.1 Hardness and Conductivity Trends........................................................... 148
4.5.2.2 Strengthening and SHR Data.................................................................... 150
4.5.3 Compression Testing Data .............................................................................. 153
4.5.4 Inhibition of GP Zone Formation.................................................................... 156
4.6 Conclusions ........................................................................................................... 165
4.7 Future Work........................................................................................................... 166
CHAPTER 5. IMPROVED STRENGTH THROUGH SOLUTE ENHANCED STRAIN
HARDENING................................................................................................................. 167

5.1 Solute Enhanced Strain Hardening........................................................................ 168
5.3 Processing and Experimental Method................................................................... 172
5.3.1 Material and Specimen extraction................................................................... 173
5.3.2 Artificial Aging of AA2524 ............................................................................ 174
5.3.3 SESH of AA2524............................................................................................ 174
5.3.4 Microstructure Characterization...................................................................... 175
5.3.5 Tensile Testing................................................................................................ 176
5.4 Experimental Results............................................................................................. 176
5.4.1 Microstructure Characterization...................................................................... 177
5.4.2 Tensile Data..................................................................................................... 183
5.5 Discussion.............................................................................................................. 188
5.5.1 Microstructure Characterization...................................................................... 188
5.5.2 Tensile Results ................................................................................................ 192

vii
5.6 Conclusions ........................................................................................................... 198
5.7 Future Work........................................................................................................... 198
CHAPTER 6. EFFECT OF SOLUTE ENHANCED STRAIN HARDENING ON
TOUGHNESS................................................................................................................. 200

6.2 Processing and Experimental Methods ................................................................. 204
6.2.1 Material and Specimen Extraction.................................................................. 204
6.2.2 Fracture Toughness Testing ............................................................................ 205
6.2.3 Fractographic Analysis.................................................................................... 220
6.3 Experimental Results............................................................................................. 221
6.3.1 Fracture Toughness Data................................................................................. 221
6.3.2 Fractography.................................................................................................... 224
6.4 Discussion.............................................................................................................. 230
6.4.1 Fracture Toughness ......................................................................................... 230
6.5 Conclusions ........................................................................................................... 234
6.6 Future Work........................................................................................................... 235
CHAPTER 7. PROJECT SUMMARY........................................................................... 237

LIST OF REFERENCES................................................................................................ 239

APPENDIX A. Compression Stress-Strain Curves........................................................ 248

APPENDIX B. Deformation Al Alloys & the Inhibition of GP Zone Formation.......... 250

APPENDIX C. Preliminary Experimental Results......................................................... 258

APPENDIX D. Characterization of Texture Produced by SPD During ECAP.............. 266

APPENDIX E. Representative Optical Micrographs ..................................................... 273

APPENDIX F. Representative Fractographs.................................................................. 290

VITA............................................................................................................................... 307

viii
List of Tables

Table 2-1: Comparison of theoretical ( ), experimental ( ), and the Peierls-Nabarro
frictional ( ) yield strengths for various metals. Note the small values of / [9].
th exp
o exp th
.... 12
Table 2-2: Slip systems for metal crystal structures. Not all systems listed are primary
slip systems. ...................................................................................................................... 14
Table 2-3: Various crystal defects observed in metals separated into class. .................... 15
Table 2-4: Hall-Petch parameters for various alloys [32]................................................. 26
Table 2-5: Precipitation sequence for various precipitation hardening alloys [31]. ......... 39
Table 2-6: Summary of strengthening mechanisms.......................................................... 47
Table 2-7: Subset of standard test methods available to extract toughness parameters [59].
........................................................................................................................................... 53
Table 2-8: Summary of the general features of plane-strain and plane-stress fracture. ... 65
Table 3-1: Alloy series with majority allowing elements. ................................................ 81
Table 3-2: Basic temper designation and descriptions. .................................................... 81
Table 3-3: Chemical composition by weight% for original alloy 2024 and higher purity
version 2524...................................................................................................................... 85
Table 4-1: Natural aging data from Figure 4-10 and Figure 4-11. ................................. 112
Table 4-2: Data from Figure 4-22 in tabular form. ......................................................... 125
Table 4-3: Values and variables used in equations 4.4 and 4.5 to construct Figure 4-34
[128]. ............................................................................................................................... 139
Table 4-4: Mean vacancy concentration (equ. 4.6) and spacing at the time of deformation
(equ. 4.7). ........................................................................................................................ 159
Table 4-5: The data plotted in Figure 4-40. .................................................................... 161
Table 4-6: Variables used to estimate the dislocation spacing. ...................................... 161
Table 5-1: Values of solute strengthening parameters for alloying additions [6]........... 170
Table 5-2: Processing parameters for the aged coupon size plates................................. 174

ix
Table 5-3: Processing parameters for the SESH coupon size plates. ............................. 175
Table 5-4: Grain sizes for each processing condition. Data graphically displayed in
Figure 5-7, Figure 5-8, and Figure 5-9. .......................................................................... 178
Table 5-5: Average hardness value and standard deviation for each processing condition.
......................................................................................................................................... 182
Table 5-6: Average tensile properties for all eight processing conditions...................... 185
Table 5-7: Partitioning coefficients (k) for various aluminum binary alloy systems [137].
......................................................................................................................................... 191
Table 5-8: Tensile properties for the peakaged material (T6-temper) and the SESH
material processed in the W-temper to a CR strain of 0.17. ........................................... 196
Table 6-1: Nominal dimensions of the compact tension specimen. ............................... 205
Table 6-2: Non-linear fracture toughness testing parameters. Specimen specific fields
have been left blank. ....................................................................................................... 211
Table 6-3: Average toughness and tensile values for the aged and SESH material. ...... 221
Table 6-4: Fractography observations for each processing condition. ........................... 225
Table 6-5: Average elemental composition of the fractured surfaces, bulk material, and
the range specified for AA2524. ..................................................................................... 229

260 Table C- 1: ECAP strain and yield strength data plotted in Figure C-4 in tabular form.
Table C-2: Yield strength values in the annealed or T4 condition for the alloys processed
and those achieved through ECAP processing. The typical range represents the maximum
and minimum values reported by others using traditional processing techniques.......... 263
Table C-3: Tensile and toughness values for AA2124................................................... 265



x
List of Figures

Figure 1-1: Density plotted against metallic element showing Aluminum, Beryllium, and
Magnesium with the lowest densities [90].......................................................................... 2
Figure 1-2: Galvanic series with materials becoming more susceptible to corrosion from
left to right [2]. .................................................................................................................... 3
Figure 1-3: A 2xxx series Al alloy with strengthening precipitates within the grain and
embrittleing grain boundary particles between grains. ....................................................... 5
Figure 1-4: Normalized fracture toughness plotted against normalized yield strength to
compare dissimilar alloys [7].............................................................................................. 6
Figure 2-1: Typical engineering stress-strain curve for a metal specimen loaded in
uniaxial tension with a broken line delineating the elastic and plastic regions. ................. 9
Figure 2-2: (a) Schematic representation of adjacent atom planes sliding past each other
and (b) the lattice energy and stress associated with the displacement. ........................... 10
Figure 2-3: Schematic representation of a dislocation moving through a crystal lattice.. 13
Figure 2-4: Face centered cubic crystal structure (a) atomic arrangement and (b)
sectioned to show a close packed plane (111) and three of the close packed directions
<110> [15]. ....................................................................................................................... 14
Figure 2-5: Resolved shear stress plotted against dislocation density for single and
polycrystalline Copper [16]. ............................................................................................. 18
Figure 2-6: Schematic representation of a shear stress-strain curve for a single crystal
[20]. ................................................................................................................................... 19
Figure 2-7: EBSD image of polycrystalline Al showing grains of various orientations
distinguished by color and separated by interfaces (grain boundaries). ........................... 22
Figure 2-8: Schematic representation of (a) a polycrystalline structure being pulled in
tension, (b) the deformation that would occur if each grain in the aggregate was
unconstrained by its neighbors and the subsequent voids and overlaps that would result,
(c) how allowing for multiple slip would eliminate the voids and overlaps (d) resulting in
a higher strain (dislocation density) near the grain boundaries [25]................................. 23
Figure 2-9: Hall-Petch plot for various alloys showing a linear relationship between yield
stress and the reciprocal square of grain size, with a slope of k [31].
y
............................. 26
Figure 2-10: Hall-Petch plot showing a change in slope (k ) at small grain sizes. Note
that the vertical axes is Vickers hardness rather than yield stress [35].
y
............................ 27

xi
Figure 2-11: Schematic representation of (a) an annealed polycrystalline structure and (b)
the resulting local plastic flow in the grain boundary region (microyielding) at the start of
deformation followed by (c) the formation of a hardened region with an increased
dislocation density that effectively reinforces the microstructure [40]. ........................... 28
Figure 2-12: Schematic representation of a solute atom (a) producing a negative
(compressive) stress field, and (b) a positive (tensile) stress field [41]............................ 30
Figure 2-13: Schematic representation of (a) an edge dislocation exerting a negative
stress field above the slip plane and positive below, (b) smaller substitutional solute
atoms with a positive stress field interacting with an edge dislocation to mutually relax
their fields, and (c) larger substitutional solute atoms with a negative stress field
interacting with an edge dislocation to mutually relax their fields [41]. .......................... 31
Figure 2-14: Critical resolved shear stress of single crystals tested at room temperature
(plateau stress) plotted against solute concentration to the 2/3 power [48]. .................... 33
Figure 2-15: The slope of each binary alloy data set shown in Figure 2 14 plotted against
the Labusch combined interaction parameter [48]............................................................ 34
Figure 2-16: Tensile yield strength plotted against cold rolling strain showing both an
increased yield point and strain hardening rate for alloys of increase levels of solute [5].
........................................................................................................................................... 35
Figure 2-17: Increased flow stress varying with the square root of solute composition for
(a) substitutional atoms and (b) interstitial atoms [9]. ...................................................... 37
Figure 2-18: Al-Cu phase diagram showing a possible solution heat treat and quench to
produce a supersaturated solid solution [41]. ................................................................... 38
Figure 2-19: Schematic representation of the molar free energy diagram for the Al-Cu
system [52]........................................................................................................................ 39
Figure 2-20: Schematic representation of (a) a dislocation approaching a spherical
precipitate that lies on its slip plane and (b&c) shearing it as it passes increasing the
precipitate-matrix interface which increases the energy of the system [9]....................... 40
Figure 2-21: Characteristic aging curves for an Al-Cu alloy [41]. ................................... 41
Figure 2-22: Schematic representation of Orowan bowing showing (a) a dislocation
approaching an array of precipitates, (b&c) bowing around them, and (d) by-passing them
leaving a dislocation loop. ................................................................................................ 43
Figure 2-23: Shear stress plotted against precipitate radius for both the shearing process
and the by-pass process, showing a switching between mechanisms at some critical
radius [9]. .......................................................................................................................... 44

xii
Figure 2-24: (a) Critical shear stress of Al-1.6%Cu specimen tested at room temperature
after being aged for various times at 190C. (b) Room-temperature stress-strain curves
after aging for select times at 190C. (b) SHR (at 2% strain) plotted against aging time
[55]. ................................................................................................................................... 45
Figure 2-25: (a) Spherical ' precipitates in a Ni-19%Cr-6%Al alloy aged at 750C for
540h by dislocation flow being sheared [56] and (b) dislocations bowing between
precipitates leaving loops behind in an Al-0.2%Au alloy deformed 5% after solutionizing
and aging at 200C for 60h. [57]. ..................................................................................... 46
Figure 2-26: Potential energy and applied force plotted against inter atomic spacing for
two adjacent atoms [58]. ................................................................................................... 49
Figure 2-27: Fracture sensitive approach design methodology depicting the
interdependence of toughness, applied stress, and flaw size. The toughness parameters are
a result of the microstructure of the material. ................................................................... 52
Figure 2-28: Schematic representation of an infinitely wide plate containing a center
crack subjected to a constant stress................................................................................... 55
Figure 2-29: Definition of coordinate system ahead of a crack tip [20]. .......................... 60
Figure 2-30: Schematic representation of thin (C)T sample in plane-stress and thick (C)T
sample in plane-strain. ...................................................................................................... 63
Figure 2-31: Experimentally observed dependence of stress intensity factor on specimen
thickness [69]. ................................................................................................................... 63
Figure 2-32: Schematic representation of a line contour surrounding a crack tip [20]. ... 66
Figure 2-33: Fracture toughness plotted against yield strength showing an inverse
relationship [74]. The limit lines represent the loci of combinations of maximum values
of K and yield strengths.
IC
................................................................................................ 68
Figure 2-34: Predicted elastic stress near a crack tip and estimation of the actual elastic-
plastic stress distribution [58]. .......................................................................................... 70
Figure 2-35: Separation of fracture types and schematic representation of each [58, 76].72
Figure 2-36: Tensile engineering stress-strain curve depicting brittle and ductile behavior.
........................................................................................................................................... 73
Figure 2-37: Schematic process of ductile failure resulting in a fractured surface that
exhibits extensive plastic deformation and dimpled rupture [58]..................................... 74
Figure 2-38: Effect of constituent particles on 2XXX and 7XXX series Al alloys [77]. . 75

xiii
Figure 2-39: Samples orientation relative to plate orientation [78]. ................................. 75
Figure 2-40: Inclusion spacing plotted against dimple size for age hardening Al alloys
[79]. ................................................................................................................................... 77
Figure 2-41: Micrograph of a cleavage fracture showing a facetted surface and no
ductility. ............................................................................................................................ 78
Figure 2-42: Micrograph of a intergranular brittle fracture showing a facetted surface and
grain structure. .................................................................................................................. 78
Figure 3-1: Tensile yield strength plotted against product thickness for 7085-T7452 and
7050-T7452 S-Basis. MMPDS MIN for 7085 is a tentative A-Basis [82]....................... 83
Figure 3-2: Ultimate tensile strength plotted against product thickness for 7085-T7452
and 7050-T7452 S-Basis. MMPDS MIN for 7085 is a tentative A-Basis [82]. ............... 83
Figure 3-3: Fracture toughness in the L-T orientation plotted against product thickness for
7085-T7452. S-Basis minimums are tentative [82]. ......................................................... 84
Figure 4-1: Tensile yield strength plotted against rolling strain for Al alloys with
increasing amounts of solute [5]. Each data set has been fitted to an exponential equation.
........................................................................................................................................... 89
Figure 4-2: Stain Hardening Rate plotted against rolling strain for Al alloys with
increasing amounts of solute [5]. Each data set has been fitted to an exponential equation.
........................................................................................................................................... 90
Figure 4-3: Schematic representation of the non-uniform cooling that results when a
work-piece is quenched into room temperature water from SHT temperature and the
subsequent residual stress from the non-uniform contraction. ......................................... 99
Figure 4-4: Effect of pre-strain followed by followed by a 90 change in direction of
uniaxial tension on the relationship of SHR and flow stress plotted against total effective
strain for (a) CP Aluminum and (b) an Al-Cu-Mg alloy [105]....................................... 102
Figure 4-5: Flow chart depicting the processing steps involved in solution heat treating
(SHT) and the uphill quenching procedure..................................................................... 106
Figure 4-6: Temperature profile associated with the processing steps depicted in Figure
4-5. .................................................................................................................................. 106
Figure 4-7: Flow chart depicting the overall experimental procedure with the deformation
process branched into three paths representing different amounts of GP zone formation.
......................................................................................................................................... 108

xiv
Figure 4-8: Tooling using for compression testing......................................................... 109
Figure 4-9: AA2524 plate that was solution heat treated and water quenched exhibiting
mid-plane cracking after light cold rolling. .................................................................... 110
Figure 4-10: Rockwell B hardness plotted against time at room temperature after
solutionizing and quenching into water (blue) and a polymer quenchent (green).......... 112
Figure 4-11: Eddy current electrical conductivity plotted against time at room
temperature after solutionizing and quenching into water (blue) and a polymer quenchent
(green). ............................................................................................................................ 113
Figure 4-12: Maximum rolling strain applied before mid-plane cracking (material CR to
strains exceeding 1.2 did not crack) and estimated time to reach 100C plotted against the
percent polymer mixed into room temperature water..................................................... 114
Figure 4-13: Rockwell B Hardness and Eddy Current Conductivity plotted against the
percent polymer mixed into room temperature water..................................................... 115
Figure 4-14: Flow chart depicting the experimental procedure used to isolate the two
possible causes of the cracking during cold rolling. ....................................................... 116
Figure 4-15: AA2524 plate that was successfully cold rolled after being solution heat
treated and water quenched followed by an additional solutionizing and polymer quench.
......................................................................................................................................... 117
Figure 4-16: Rockwell B Hardness plotted against natural aging time (time at room
temperature after quenching) for the samples cold rolled in the half hardened condition
(1.3hr).............................................................................................................................. 119
Figure 4-17: Eddy Current Electrical Conductivity plotted against natural aging time
(time at room temperature after quenching) for the samples cold rolled in the half
hardened condition (1.3hr).............................................................................................. 119
Figure 4-18: Rockwell B Hardness plotted against Rolling Strain with the data for each
temper fitted to a linear trend line................................................................................... 121
Figure 4-19: The Hardness Increment resulting from dislocations only plotted against
Rolling Strain with data for each temper fitted to a linear trend line. ............................ 122
Figure 4-20: Eddy Current Electrical Conductivity Drop upon rolling plotted against
rolling strain. ................................................................................................................... 123
Figure 4-21: The Hardness Increment resulting from dislocations only plotted against
eddy current electrical conductivity drop upon rolling................................................... 124

xv
Figure 4-22: Compressive yield strength plotted against cold rolling strain for the each of
the three tempers. ............................................................................................................ 126
Figure 4-23: Strain Hardening Rate (SHR) plotted against the Cold Rolling (CR) flow
stress minus the Yield Stress........................................................................................... 127
Figure 4-24: Compressive yield strength increment plotted against the conductivity drop
upon rolling. .................................................................................................................... 128
Figure 4-25: Compression stress-strain curves for the non-rolled material tested in the W,
1/2H, and T4 temper. ...................................................................................................... 129
Figure 4-26: SHR plotted against compressive strain for the non-rolled material tested in
the W, 1/2H, and T4 temper............................................................................................ 130
Figure 4-27: SHR plotted against flow stress with the yield stress subtracted for the non-
rolled material tested in the W, 1/2H, and T4 temper..................................................... 131
Figure 4-28: SHR plotted against total strain (rolling strain plus compressive strain) for
the material cold rolled in the T4-temper. ...................................................................... 133
Figure 4-29: SHR plotted against flow stress with the yield stress subtracted for the
material cold rolled in the T4-temper. ............................................................................ 133
Figure 4-30: SHR plotted against total strain (rolling strain plus compressive strain) for
the material cold rolled in the 1/2H-temper. ................................................................... 134
Figure 4-31: SHR plotted against flow stress with the yield stress subtracted for the
material cold rolled in the 1/2H-temper.......................................................................... 134
Figure 4-32: SHR plotted against total strain (rolling strain plus compressive strain) for
the material cold rolled in the W-temper. ....................................................................... 135
Figure 4-33: SHR plotted against flow stress with the yield stress subtracted for the
material cold rolled in the W-temper. ............................................................................. 135
Figure 4-34: Diffusion Coefficient plotted against temperature for published high
temperate data extrapolated down to temperatures that natural aging occurs at and the
experimentally estimated values for different quenching rates. ..................................... 140
Figure 4-35: Vacancy concentration plotted against temperature showing the excess
quenched-in vacancies present after SHT....................................................................... 142
Figure 4-36: Shape change (curling) of water quenched material during an EDM surface
pass. The normal to surface of the work piece faces toward the center of the curling
indicating a residual tensile surface stress. ..................................................................... 146

xvi
Figure 4-37: Bar graph comparing the strain hardening exponent (n) in cold rolling and
compression relative to the material temper. .................................................................. 154
Figure 4-38: Bar graph comparing the strain hardening parameter (K) in cold rolling and
compression relative to the material temper. .................................................................. 155
Figure 4-39: Mean vacancy diffusion length plotted against time at room temperature for
a ten minute interval (equ. 4.8). ...................................................................................... 159
Figure 4-40: Feature spacing and Flow Stress plotted against Dislocation density. Note
that the two Y-axes can only be read against the X-axes and not each other. ................ 161
Figure 4-41: Dislocation spacing for hard (solid lines) and soft (broken lines) obstacles
plotted against rolling strain for the data in Figure 24.................................................... 164
Figure 5-1: Typical strength of Al alloys [132]. ............................................................. 168
Figure 5-2: Delta K (K K ) plotted against atomic fraction of solute [5].
alloy Al
............ 169
Figure 5-3: Yield Strength plotted against X-Ray line Breadth for Aluminum samples
with various amounts of solute for the alloying elements listed in Table 2 [6].............. 171
Figure 5-4: Flow chart depicting the experimental procedure for the SEHS part of the
current study.................................................................................................................... 172
Figure 5-5: Schematic representation of a coupon size plate with tensile and C(T)
specimen nested inside so that the loading direction is parallel to the rolling direction. 173
Figure 5-6: Schematic showing the dimensions of a the ASTM sub-size tensile bars
extracted from the processed plates. ............................................................................... 176
Figure 5-7: Grain size for each of the four aged microstructures and the average of the
four in the RD, TD, and ND directions........................................................................... 178
Figure 5-8: Grain size for each of the four CR microstructures in the RD, TD, and ND
directions and the average of the aged grain sizes. ......................................................... 179
Figure 5-9: Grain aspect ratio (RD/ND) for each processing condition. ........................ 180
Figure 5-10: Vickers microhardness profiles for the aged plates taken on the RD plane in
the ND direction.............................................................................................................. 181
Figure 5-11: Vickers microhardness profiles for the SESH plates taken on the RD plane
in the ND direction.......................................................................................................... 182

xvii
Figure 5-12: EDS through thickness elimental composition profile for the material in the
T4-tmeper........................................................................................................................ 183
Figure 5-13: Ultimate tensile strength (UTS) plotted against yield strength for all eight
processing conditions...................................................................................................... 184
Figure 5-14: Engineering stress plotted against engineering strain. ............................... 185
Figure 5-15: Percent elongation plotted against yield strength for all eight processing
conditions. ....................................................................................................................... 186
Figure 5-16: Reduction in area plotted against yield strength for all eight processing
conditions. ....................................................................................................................... 187
Figure 5-17: Area under the engineering stress-strain curve plotted against yield strength.
......................................................................................................................................... 188
Figure 5-18: Stress plotted against aging time at 190C for AA2024 and AA2524. Solid
lines are the aging response of AA2024. Broken lines are trend lines through the
experimentally measured AA2524 values. ..................................................................... 192
Figure 5-19: Ultimate tensile strength plotted against yield strength for 2xxx series alloys.
......................................................................................................................................... 193
Figure 5-20: True stress plotted against true strain. Solid lines represent continuous
compression testing data. Compressive yield strength-CR strain is represented by discrete
solid data points with broken trend lines. The equations for the trend lines are given.
Tensile yield strength-CR strain is represented by hollow data points and has not been
fitted to trend lines. ......................................................................................................... 195
Figure 5-21: Strain hardening exponent (n) plotted against tensile yield strength for
AA2524 in all eight processing conditions. .................................................................... 197
Figure 6-1: Yield strength plotted against fracture toughness for wrought precipitation
harden aluminum alloys [132]. ....................................................................................... 201
Figure 6-2: Variation in slow bend Charpy toughness with yield strength for various heat
treatments [76]. ............................................................................................................... 202
Figure 6-3: Variation of Charpy toughness with areal fraction of grain boundary , at
constant matrix yield strength [76]. ................................................................................ 203
Figure 6-4: Schematic of a compact tension specimen with nominal dimensions (units of
millimeters). Specimen is in the L-T orientation. ........................................................... 205
Figure 6-5: Side-grooved compact tension specimen with dry graphite lubricant. ........ 207

xviii
Figure 6-6: Compact tension specimen showing net thickness and side-groove angle. . 207
Figure 6-7: Load plotted against COD displacement for the unloading used to update the
crack length (T6 specimen). Note that the loading data is not plotted. .......................... 210
Figure 6-8: Half of a fractured C(T) specimen showing (left to right) the knife edge on
the load line that the COD was mounted on, EDM starter notch, fatigue pre-crack, ductile
tearing from testing, fatigue marker region, and the mechanical overloading when the
critical crack length was reached. ................................................................................... 212
Figure 6-9: C(T) specimen width plotted against the net thickness showing a serious of
nine measurements for the load line, EDN starter crack, fatigue pre-crack front, and the
end of the ductile tearing (beginning of the fatigue marker). Note that the first and last
measurements are 0.005W in from the root radius of the side-grooves. ........................ 213
Figure 6-10: Graphical construction depicting the determination of J [139].
Q
.............. 218
Figure 6-11: Critical nonlinear energy release rate plotted against yield strength for the
aged and SESH material. ................................................................................................ 222
Figure 6-12: Fracture toughness calculated from critical nonlinear energy release rate
plotted against yield strength. ......................................................................................... 222
Figure 6-13: Crack extension plotted against energy release rate for both the aged and
SESH material................................................................................................................. 223
Figure 6-14: Tearing modulus plotted against yield strength for both the aged and SESH
material. .......................................................................................................................... 224
Figure 6-15: SEM fractograph looking at the RD plane of the material in the T6-temper at
100x showing a brittle appearance.................................................................................. 226
Figure 6-16: SEM fractograph looking at the RD plane of the material in the T4-temper at
1,000x showing GBDF. .................................................................................................. 226
Figure 6-17: SEM fractograph looking at the RD plane of the material SESH in the W-
temper to a true strain of 0.17 at 100x showing a ductile appearance. ........................... 227
Figure 6-18: SEM fractograph looking at the RD plane of the material SESH in the W-
temper to a true strain of 0.17 at 1,000x showing intergranular dimpled rupture. ......... 227
Figure 6-19: Approximate dimple spacing plotted against fracture toughness for each
processing condition. ...................................................................................................... 228
Figure 6-20: Elemental compostion of the fratured surfaces for each processing condition.
......................................................................................................................................... 229

xix
Figure 6-21: Fracture toughness plotted against yield strength for 2524 and 2xxx series
alloys. .............................................................................................................................. 232

Figure A-1: True stress plotted against true strain for cylindrical compression specimen
tested after being cold rolled to various strains in the W-temper. .................................. 248
Figure A-2: True stress plotted against true strain for cylindrical compression specimen
tested after being cold rolled to various strains in the 1/2H-temper. .............................. 249
Figure A- 3: True stress plotted against true strain for cylindrical compression specimen
tested after being cold rolled to various strains in the T4-temper................................... 249

.............................................................................................................................. 259
Figure C-1: Image of ECAP fixture placed in an Instron 58R1127 screw driven load
frame.
Figure C-2: Image of a billet as it is (a) passing through the shear plane, (b) almost
through the shear plane, and (c) pushed past the shear plane by the next billet. ............ 259
Figure C-3: True compressive stress plotted against ECAP strain for commercial purity
P0815. ............................................................................................................................. 260
Figure C-4: True compressive stress plotted against true compressive strain post ECAP
processing. ...................................................................................................................... 261
Figure C- 5: {001}, {011}, and {111} pole figures for the fully annealed P0815 revealing
a strong {112}<111> single texture component. The ND direction corresponds to long
direction of the billet. ...................................................................................................... 262
Figure C- 6: {001}, {011}, and {111} pole figures for the P0815 processed to a strain of
10 revealing a {411}<110> shear texture. The ND direction corresponds to long direction
of the billet. ..................................................................................................................... 262
Figure C-7: Effect of a single pass through the ECAP fixture on AA2524 if quenched into
water or polymer after SHT. ........................................................................................... 263
Figure C-8: Compressive stress-strain curves for AA2524, AA5182, and P0815 after
ECAP processing. ........................................................................................................... 264


................................................................................................. 274
Figure E-1: Micrograph of the TD plane of the T4 material etched with Keller's reagent at
a magnification of 100x.
Figure E-2: Micrograph of the TD plane of the T4 material etched with Keller's reagent at
a magnification of 500x. ................................................................................................. 274

xx
Figure E-3: Micrograph of the RD plane of the T4 material etched with Keller's reagent at
a magnification of 100x. ................................................................................................. 275
Figure E-4: Micrograph of the RD plane of the T4 material etched with Keller's reagent at
a magnification of 500x. ................................................................................................. 275
Figure E-5: Micrograph of the TD plane of the UA material etched with Keller's reagent
at a magnification of 100x. ............................................................................................. 276
Figure E-6: Micrograph of the TD plane of the UA material etched with Keller's reagent
at a magnification of 500x. ............................................................................................. 276
Figure E-7: Micrograph of the RD plane of the UA material etched with Keller's reagent
at a magnification of 100x. ............................................................................................. 277
Figure E-8: Micrograph of the RD plane of the UA material etched with Keller's reagent
at a magnification of 500x. ............................................................................................. 277
Figure E-9: Micrograph of the TD plane of the T6 material etched with Keller's reagent at
a magnification of 100x. ................................................................................................. 278
Figure E-10: Micrograph of the TD plane of the T6 material etched with Keller's reagent
at a magnification of 500x. ............................................................................................. 278
Figure E-11: Micrograph of the RD plane of the T6 material etched with Keller's reagent
at a magnification of 100x. ............................................................................................. 279
Figure E-12: Micrograph of the RD plane of the T6 material etched with Keller's reagent
at a magnification of 500x .............................................................................................. 279
Figure E-13: Micrograph of the TD plane of the OA material etched with Keller's reagent
at a magnification of 100x. ............................................................................................. 280
Figure E-14: Micrograph of the TD plane of the OA material etched with Keller's reagent
at a magnification of 500x. ............................................................................................. 280
Figure E-15: Micrograph of the RD plane of the OA material etched with Keller's reagent
at a magnification of 100x. ............................................................................................. 281
Figure E-16: Micrograph of the RD plane of the OA material etched with Keller's reagent
at a magnification of 500x. ............................................................................................. 281
Figure E-17: Micrograph of the TD plane of the T4 material cold rolled to a true strain of
0.16 etched with Keller's reagent at a magnification of 100x......................................... 282

xxi
Figure E-18: Micrograph of the TD plane of the T4 material cold rolled to a true strain of
0.16 etched with Keller's reagent at a magnification of 500x......................................... 282
Figure E-19: Micrograph of the RD plane of the T4 material cold rolled to a true strain of
0.16 etched with Keller's reagent at a magnification of 100x......................................... 283
Figure E-20: Micrograph of the RD plane of the T4 material cold rolled to a true strain of
0.16 etched with Keller's reagent at a magnification of 500x......................................... 283
Figure E-21: Micrograph of the TD plane of the W material cold rolled to a true strain of
0.17 etched with Keller's reagent at a magnification of 100x......................................... 284
Figure E-22: Micrograph of the TD plane of the W material cold rolled to a true strain of
0.17 etched with Keller's reagent at a magnification of 500x......................................... 284
Figure E-23: Micrograph of the RD plane of the W material cold rolled to a true strain of
0.17 etched with Keller's reagent at a magnification of 100x......................................... 285
Figure E-24: Micrograph of the RD plane of the W material cold rolled to a true strain of
0.17 etched with Keller's reagent at a magnification of 500x......................................... 285
Figure E-25: Micrograph of the TD plane of the W material cold rolled to a true strain of
0.31 etched with Keller's reagent at a magnification of 100x......................................... 286
Figure E-26: Micrograph of the TD plane of the W material cold rolled to a true strain of
0.31 etched with Keller's reagent at a magnification of 500x......................................... 286
Figure E-27: Micrograph of the RD plane of the W material cold rolled to a true strain of
0.31 etched with Keller's reagent at a magnification of 100x......................................... 287
Figure E-28: Micrograph of the RD plane of the W material cold rolled to a true strain of
0.31 etched with Keller's reagent at a magnification of 500x......................................... 287
Figure E-29: Micrograph of the TD plane of the T4 material cold rolled to a true strain of
0.41 etched with Keller's reagent at a magnification of 100x......................................... 288
Figure E-30: Micrograph of the TD plane of the T4 material cold rolled to a true strain of
0.41 etched with Keller's reagent at a magnification of 500x......................................... 288
Figure E-31: Micrograph of the RD plane of the T4 material cold rolled to a true strain of
0.41 etched with Keller's reagent at a magnification of 100x......................................... 289
Figure E-32: Micrograph of the RD plane of the T4 material cold rolled to a true strain of
0.41 etched with Keller's reagent at a magnification of 500x......................................... 289


xxii
Figure F-1: Fractograph of the RD plane of the material in the T4-temper at
magnification of 100x. .................................................................................................... 291
Figure F-2: Fractograph of the RD plane of the material in the T4-temper at
magnification of 500x. .................................................................................................... 291
Figure F-3: Fractograph of the RD plane of the material in the T4-temper at
magnification of 1,000x. ................................................................................................. 292
Figure F-4: Fractograph of the RD plane of the material in the T4-temper at
magnification of 2,000x. ................................................................................................. 292
Figure F-5: Fractograph of the RD plane of the material in the UA condition at
magnification of 100x. .................................................................................................... 293
Figure F-6: Fractograph of the RD plane of the material in the UA condition at
magnification of 500x. .................................................................................................... 293
Figure F-7: Fractograph of the RD plane of the material in the UA condition at
magnification of 1,000x. ................................................................................................. 294
Figure F-8: Fractograph of the RD plane of the material in the UA condition at
magnification of 2,000x. ................................................................................................. 294
Figure F-9: Fractograph of the RD plane of the material in the T6-temper at
magnification of 100x. .................................................................................................... 295
Figure F-10: Fractograph of the RD plane of the material in the T6-temper at
magnification of 500x. .................................................................................................... 295
Figure F-11: Fractograph of the RD plane of the material in the T6-temper at
magnification of 1,000x. ................................................................................................. 296
Figure F-12: Fractograph of the RD plane of the material in the T6-temper at
magnification of 2,000x. ................................................................................................. 296
Figure F-13: Fractograph of the RD plane of the material in the OA condition at
magnification of 100x. .................................................................................................... 297
Figure F-14: Fractograph of the RD plane of the material in the OA condition at
magnification of 500x. .................................................................................................... 297
Figure F-15: Fractograph of the RD plane of the material in the OA condition at
magnification of 1,000x. ................................................................................................. 298

xxiii
Figure F-16: Fractograph of the RD plane of the material in the OA condition at
magnification of 2,000x. ................................................................................................. 298
Figure F-17: Fractograph of the RD plane of the T4 material cold rolled to a true strain of
0.16 at magnification of 100x. ........................................................................................ 299
Figure F-18: Fractograph of the RD plane of the T4 material cold rolled to a true strain of
0.16 at magnification of 500x. ........................................................................................ 299
Figure F-19: Fractograph of the RD plane of the T4 material cold rolled to a true strain of
0.16 at magnification of 1,000x. ..................................................................................... 300
Figure F- 20: Fractograph of the RD plane of the T4 material cold rolled to a true strain
of 0.16 at magnification of 2,000x.................................................................................. 300
Figure F- 21: Fractograph of the RD plane of the W material cold rolled to a true strain of
0.17 at magnification of 100x. ........................................................................................ 301
Figure F-22: Fractograph of the RD plane of the W material cold rolled to a true strain of
0.17 at magnification of 500x. ........................................................................................ 301
Figure F-23: Fractograph of the RD plane of the W material cold rolled to a true strain of
0.17 at magnification of 1,000x. ..................................................................................... 302
Figure F-24: Fractograph of the RD plane of the W material cold rolled to a true strain of
0.17 at magnification of 2,000x. ..................................................................................... 302
Figure F-25: Fractograph of the RD plane of the W material cold rolled to a true strain of
0.31 at magnification of 100x. ........................................................................................ 303
Figure F-26: Fractograph of the RD plane of the W material cold rolled to a true strain of
0.31 at magnification of 500x. ........................................................................................ 303
Figure F-27: Fractograph of the RD plane of the W material cold rolled to a true strain of
0.31 at magnification of 1,000x. ..................................................................................... 304
Figure F-28: Fractograph of the RD plane of the W material cold rolled to a true strain of
0.31 at magnification of 2,000x. ..................................................................................... 304
Figure F-29: Fractograph of the RD plane of the T4 material cold rolled to a true strain of
0.41 at magnification of 100x. ........................................................................................ 305
Figure F-30: Fractograph of the RD plane of the T4 material cold rolled to a true strain of
0.41 at magnification of 500x. ........................................................................................ 305

xxiv
Figure F-31: Fractograph of the RD plane of the T4 material cold rolled to a true strain of
0.41 at magnification of 1,000x. ..................................................................................... 306
Figure F-32: Fractograph of the RD plane of the T4 material cold rolled to a true strain of
0.41 at magnification of 2,000x. ..................................................................................... 306


xxv
Abstract

Solute Enhanced Strain Hardening of Aluminum Alloys
to Achieve Improved Combinations of Strength and Toughness

Christopher J. Hovanec
1,2

Roger D. Doherty, D. Phil.
1

Surya R. Kalidindi, Ph.D.
3


1
Department of Materials Science and Engineering, Drexel University
2
Naval Air Warfare Center - Aircraft Division, Patuxent River NAS
3
Department of Mechanical Engineering and Mechanics, Drexel University




The feasibility of achieving improved combinations of strength and toughness in
aluminum alloy 2524 through solute enhanced strain hardening (SESH) has been
explored in this study and shown to be viable. The effectiveness of SESH is directly
dependent on the strain hardening rate (SHR) of the material being processed. Aluminum
alloy 2524 naturally ages to the T4-temper after solution heat treating and quenching. The
SHR of strain free and post cold rolled material as a function of natural aging time has
been measured by means of simple compression. It has been determined that the SHR of
AA2524 is more effective with solute in solution rather than clustered into GP zones. It
has also been shown that the typical rapid formation of GP zones at room temperature
(natural aging) is inhibited by moderate cold rolling strains (
CR
0.2) through
dislocation aided vacancy annihilation.
The practical limitations of quenching rate have been determined using hardness
and eddy current electrical conductivity measurements. It has been shown that too slow
of a quench rate results in solute being lost to both the formation of GP zones and
embrittling precipitates during the quench, while too rapid of a quench rate results in

xxvi
mid-plane cracking of the work piece during the SESH processing. The mid-plane
cracking was overcome by using an uphill quenching procedure to relieve residual
stresses within the work piece. Aluminum alloy 2524 strengthened through SESH to a
yield strength 11% greater than that in the T6-Temper exhibits: equivalent toughness, 5%
greater UTS, 1% greater elongation, 7% greater R.A., and absorbs 15% more energy
during tensile testing. At yield strengths comparable to published data for 2x24 alloys,
the SESH 2524 exhibited up to a 60% increase in fracture toughness. The fractured
surfaces of the SESH material exhibited transgranular dimpled rupture as opposed to the
grain boundary ductile fracture (GBPF) observed in the artificially aged material.


xxvii













Page intentionally left black.

























1
CHAPTER 1. INTRODUCTION

There is an ever present drive to exploit the interdependent processing-
microstructure-properties relationship to enhance the performance of a material. The
primary reason for this is that high-end design is often limited by material performance.
Some of the most common material requirements are the response to both static and
dynamic loading, environmental susceptibility (corrosion), and thermal stability. An
additional challenge often faced by a design team is the need for increased performance
coupled with increased efficiency. It is almost universally true that both of these
objectives are achieved by reducing weight. This is particularly true in the transportation
field, especially the aerospace industry. The project manager for the Airbus A340
estimates that a 5% reduction in weight would translate into of savings of over $6 million
in transatlantic travel annually per aircraft [88]. The automotive industry estimates that a
10% reduction in weight would improve fuel efficiency by 6% to 8% [89]. Being that
density is intrinsic to a material, weight is not a material variable that can be significantly
altered through processing techniques. This means that informed alloy selection must be
the step that precedes processing-microstructure-properties optimization. The three alloy
systems that will inherently be the lightest are those primarily composed of Aluminum
(Al), Beryllium (Be), or Magnesium (Mg) as the base metal, as shown in Figure 1-1.

2
0
5000
10000
15000
20000
25000
D
e
n
s
i
t
y

(
K
g
/
m
3
)
A
l
u
m
in
u
m
B
e
r
y
l
li
u
m
B
r
a
s
s
C
o
b
o
lt
C
o
p
p
e
r
G
o
ld
H
a
t
e
l
lo
y
I
n
c
o
n
e
l
I
r
o
n
L
e
a
d
M
a
g
n
e
s
i
u
m
N
i
c
k
e
l
S
i
l
v
e
r
P
l
a
t
i
n
u
m
S
t
e
e
l
T
i
n
T
i
t
a
n
i
u
m
T
u
n
g
s
t
e
n
Z
i
n
c
W Vg =

Figure 1-1: Density plotted against metallic element showing Aluminum, Beryllium, and
Magnesium with the lowest densities [90].



Upon more careful scrutiny, Beryllium and Magnesium can be eliminated from
the list of potential base metals, leaving Aluminum as the most desirable choice.
Beryllium is quickly eliminated due to the serious health hazards that accompany it [1].
The Cu-Be alloys currently used in bushing and bearing applications on Navy Aircraft
are in the process of being restricted for use on future Naval Aviation Systems due to the
heath risk posed to the Fleet maintainers. Magnesium is a very anodic material, as shown
in Figure 1-2 [2], which makes it environmentally susceptible. Inevitably, a mechanical
system will be composed of many components consisting of various alloys, and the active
nature of Mg based alloys will result in preferential corroding when coupled with any
dissimilar metal. This can be overcome with coating technologies however, coatings
offset the weight savings and add additional cost (although Mg may still be an attractive

3
option for some applications). With the exception of some select helicopter applications,
the high activity of Mg has resulted in it being prohibited for use on Naval aircraft. The
combination of low density and an inert oxide film therefore makes Al based alloy
systems an appealing option when trying to increase both performance and efficiency.



Figure 1-2: Galvanic series with materials becoming more susceptible to corrosion from
left to right [2].


4
To help put the weight saving of Aluminum in perspective, a typical Aluminum
part is 40% to 50% lighter then a comparable steel part [91]. There are performance
requirements met by steels that Al based alloys are unable to achieve. For example, both
commercial and military airframes are predominantly Al alloys, while their landing gear
components remain high strength steel. However, if the mechanical properties of Al
alloys are improved they may be able to displace steels in some of their current
applications. An additional benefit to improving the mechanical properties of Al alloys is
that, in applications where they are already being used, the gauge thickness necessary to
perform at a given stress level may be reduced, further reducing weight (provided that the
application isn't limited by elastic stiffness). The fundamental motivation behind the
current study is a reduction in weight through improved mechanical properties.
In mechanical design the two most critical material properties are arguably yield
strength (
y
) and fracture toughness (K
IC
). High strength Al alloys (2xxx and 7xxx)
achieve their properties through precipitation hardening. These alloys have been designed
to respond to this strengthening method, and precipitation hardening is very effective.
Precipitation hardening is also a desirable strengthening method because it allows for
components to be shaped at relatively low strengths (O-temper) before heat treating to
their service strength. Precipitation hardening can also be easily performed on
complicated shapes. An unavoidable consequence of the precipitation hardening process
is that, while the strengthening precipitates are forming in the interior of the grains, there
are also precipitates forming on the grain boundaries, as shown in Figure 1-3.

5

Figure 1-3: A 2xxx series Al alloy with strengthening precipitates within the grain and
embrittleing grain boundary particles between grains.



These grain boundary precipitates are known to degrade the toughness of the material.
This undesirable microstructure feature can be avoided if other strengthening methods are
employed. The oldest, and still commonly used, strengthening method in metals is strain
hardening (work hardening). This has been ignored as a viable strengthening route in Al
based alloys due to their high stacking fault energy which promotes dynamic recovery. It
has been demonstrated that potent strain hardening rates (d/d) can be achieved in heat
treatable Al alloys provided the solute remains in solution or clusters (GP zones) [3-6].
As the solute is precipitated out of the matrix, the strain hardening rate decreases at
equivalent flow stresses [143]. This prompts the question; can Al alloys be produced that
exhibit better combinations of strength and toughness through solute enhanced strain
hardening (SESH)?
A further indication that improved combinations of strength and toughness might
be possible was put forth by [7] while studying Cobalt based alloys. In an effort to
compare the combinations of strength and toughness achievable in different alloy

6
systems, these properties were normalized and plotted against each other, as shown in
Figure 1-4. The effect of normalizing the properties is that the data can be examined in
the absence of inherent bond strength (this is commonly done for yield strength).

0
50
100
150
200
250
300
350
0.013 0.015 0.017 0.019 0.021 0.023 0.025
Yield Strength/G
F
r
a
c
t
u
r
e

T
o
u
g
h
n
e
s
s
/
(
E

)
1
/
2

TRIP Steels (bcc)
Aluminum Alloys (fcc)
GBPs
Cobalt Alloy (fcc)
No GBPs
Q&T Steel
Maraging Steel (bcc)
Figure 1-4: Normalized fracture toughness plotted against normalized yield strength to
compare dissimilar alloys [7].


It might be expected that all face-centered cubic (fcc) and body-centered cubic
(bcc) alloys should fall on the same line. This is obviously not the case. The alloys
depicted in Figure 1-4 achieve their mechanical properties through different processing
methods, producing different microstructures that behave differently when the material is
loaded. Examples of this are the two fcc alloys, aluminum and cobalt. The data for these
two alloys is expected to lie on top of each other but it does not. The cobalt alloy is strain
hardened without any grain boundary particles (GBPs) while the aluminum alloy is
precipitation hardened with embrittling GBPs. Elimination of the GBPs by changing the
strengthening mechanism should improve the properties of the Al alloys moving them
closer to those observed for the cobalt alloy.

7
This study focuses on the processing-macrostructure-properties interrelationship
of Al alloys and how the microstructures produced by various strengthening mechanisms
affect fracture toughness. The overall objectives are to:
1) Determine the feasibility of achieving improved combinations of strength and
toughness in Al alloys by solute enhanced strain hardening (SESH).
2) Determine if strain hardening of Al alloys is more effective with solute in solution
or clustered into GP zones.
3) Investigate the formation of GP zones in the presence of a deformed
microstructure.

8
CHAPTER 2. BACKGROUND

The relevant background for this project is separated into six sections. The
discussion will start by reviewing the basic concepts behind metal deformation. This will
then be extended into how the resistance to deformation can be increased (strengthening
methods) both at the point of yielding and throughout the metal flow process (strain
hardening rate). The discussion will be expanded to include fracture and how this
dynamic process is described/characterized (linear and non-linear fracture mechanics).
Finally, the inverse relationship between strength and toughness will be described, as
well as how the various microstructures produced by different strengthening methods
effect this relationship.

2.1 Plastic Deformation


The desire to understand and describe what controls a metals yield strength and
how it plastically flows under an external force has been, and continues to be, of great
interest to the science and engineering community. Within this section, the basic
principles of yielding and plastic deformation in metallic systems will be reviewed with
an emphasis on aluminum alloys. It is assumed that the reader has a basic understanding
of these phenomena.
It is well known that under an externally applied load a metallic component will
exhibit elastic-plastic behavior with parabolic hardening, as shown in Figure 2-1.

9
0
100
200
300
400
500
600
700
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
Strain
S
t
r
e
s
s
Elastic Plastic

y

d
d
o
c
E
o
c
=

Figure 2-1: Typical engineering stress-strain curve for a metal specimen loaded in
uniaxial tension with a broken line delineating the elastic and plastic regions.


There is a wealth of information that can be extracted from a plot such as the one
depicted in Figure 2-1, although at this point in the discussion the focus will be on the
yield point (
y
) and the ensuing metal flow that follows yielding. In the early stages of
loading (elastic region) the stress-strain relationship is linear and the strain is recoverable.
Metals are Hookean solids and this proportional relationship, which defines the Youngs
modulus (E) of a material, continues until a critical point is reached. This critical point,
which is the transition between recoverable elastic strain and non-recoverable plastic
strain, is termed the yield strength (also referred to as the elastic limit or proportional
limit). The pertinent question is, what determines the amount of stress that can be
imposed on a metal before it plastically deforms?
The theoretical yield strength of a crystalline material with a perfect lattice was
calculated almost a hundred years ago by Frenkel [8]. Consider two adjacent atomic
planes sliding past each other in an idealized shearing () process, as shown in Figure

10
2-2a. There is an internal energy associated with the relative position of the atoms
throughout the shearing process, with the equilibrium energy being associated with the
positions where the repeating unit is restored. This low energy position coincides with a
displacement equal to the Burgers vector (b). Moving the atoms away from their
equilibrium requires a stress to overcome an energy barrier (E), which reaches a
maximum at half an atomic spacing. This stress is equal to the slope (first derivative) of
the energy curve which is the stress required to produce the displacement. The lattice
energy and associated shear stress is shown in Figure 2-2b.


(a)








-1
4.4
Displacement (x)
E
n
e
r
g
y

/

S
t
r
e
s
s
Latice Energy
Shear Stress


b
b
2

a
(b)
E

max

Figure 2-2: (a) Schematic representation of adjacent atom planes sliding past each other
and (b) the lattice energy and stress associated with the displacement.


11
For simplicity, it is assumed that the shear stress can be represented as a sine function of
the form:

max max
2
sin
2 x x
b b
t t
t t t = ~ (2.1)
where x is the displacement, which can be simplified using the small angle
approximation. It is also known that the elastic stress-strain relationship follows Hookes
law:
G t = (2.2)
where G is the shear modulus and is the shear strain. Once again, due to the scale of the
problem, an approximation can be made that the shear strain is:

x
a
= (2.3)
with a being the distance between planes. Substituting equations (2.2) and (2.3) into (2.1)
and rearranging to solve for the maximum theoretical shear strength of the crystal
produces:

max
2 2
Gb G
a
t
t t
= ~ (2.4)
which can be further simplified due to the Burgers vector and planar spacing being
approximately the same.
The theoretical shear stresses predicted by equation (2.4) are orders of magnitude
larger than those experimentally observed. This was immediately recognized and the
validity of approximating the stress-displacement curve as a sine wave was questioned.
Further calculations, using more realistic nonsymmetrical functions, to more accurately
represent the stress-displacement relationship were preformed which predicted lower

12
yield strengths. The refined analysis predicts the theoretical yield strength to be much
closer to:

max
30
G
t ~ (2.5)
which is still an overestimate, as shown in Table 2-1 [9].


Table 2-1: Comparison of theoretical (
th
), experimental (
exp
), and the Peierls-Nabarro
frictional (
o
) yield strengths for various metals. Note the small values of
exp
/
th
[9].

Material

th
(G/30)
(MPa)

exp

(MPa)

exp
/
th

o
(MPa)
Ag 1000 0.37 0.00037 ~1.6E-23
Al 900 0.78 0.00087 ~1.4E-23
Cu 1400 0.49 0.00035 ~2.2E-23
Ni 2600 3.20 0.00123 ~4.2E-23
-Fe 2600 27.50 0.01058 ~4.2E-23


The result of these calculations and the large discrepancy between the theoretical values
and the experimentally measured single crystal values lead to the conclusion that metals
must not deform by entire planes of atoms shearing past each other. By 1934, the concept
of lattice defects was put forth to explain the significantly lower stresses measured in
experiments. Taylor, Orowan, and Polanyi [10-12] all independently proposed the idea of
an extra plane of atoms inserted into a perfect crystal structure, known as a dislocation
(line defect). The significance of introducing a dislocation into a perfect lattice is that
plastic flow no longer requires as large of an increase in energy, and therefore stress,
associated with displacing an entire plane of atoms.
If an extra half plane of atoms is inserted into the lattice, as shown in Figure 2-3,
a displacement equal to the Burgers vector can be achieved by overcoming a local energy
barrier rather then the larger barrier of the entire plane. This is a result of the dislocation

13
being pushed through the structure in sequential atomic displacements rather than a
simultaneous displacement.


Figure 2-3: Schematic representation of a dislocation moving through a crystal lattice.


It should be clarified that the atoms originally present in the extra half plan of atoms are
not being pushed through the structure during metal flow, but rather the unfulfilled
metallic bond is what is flowing. The stress to move a dislocation through a perfect lattice
has a similar form as that shown in Figure 2-2, although the magnitude is significantly
less. This frictional stress (
o
) was originally described by Peierls and modified by
Nabarro [13] and is thus often referred to as the Peierls-Nabarro stress. The Peierls-
Nabarro stress is what must be overcome to activate a dislocation and produce yielding in
an ideal crystal lattice, which is given by the relationship:

2
2
(1 )
a
w
b
b
o
Ge Ge
t
t
u
t

~ ~ (2.6)
where is Poissons ratio and w is the dislocation width or distance that the dislocation
will distort the lattice (approximately 10 atomic spacings for Al, [14]). The form of
equation 2.6 given as a function of dislocation width is more appropriate for metallically
bonded materials and was used to calculate the values in Table 2-1 (w/b was
approximated to be 10 for each material). Examination of equation 2.6 produces some
significant insights into slip. First, slip will more easily occur when the distance between

14
planes (a) is large. The inter-planer spacing will be the greatest between close packed
planes. Secondly, slip will more easily occur when the slip distance (b) is smallest, which
will be in the most closely packed atomic direction. This gives rise to the fundamental
concept of the slip system, which is that slip will occur in a close packed plane and in a
close packed direction. Common slip systems for metallic crystal structures are listed in
Table 2-2 and schematically shown in Figure 2-4 for aluminum.

Table 2-2: Slip systems for metal crystal structures. Not all systems listed are primary
slip systems.
Crystal
Structure
Slip
Plane
Slip
Direction
Number
of planes
Number of
directions per
plane
Number of
slip systems
fcc {111} <110> 4 3 12
{110} <111> 6 2 12
{112} <111> 12 1 12 bcc
{123} <111> 24 1 24
{0001} <1120> 1 3 3
{1010} <1120> 3 1 3 hcp
{1011} <1120> 6 1 6



Figure 2-4: Face centered cubic crystal structure (a) atomic arrangement and (b)
sectioned to show a close packed plane (111) and three of the close packed directions
<110> [15].
(b) (a)


15
The criteria set forth for plastic yielding in this section, overcoming the frictional
stress required to activate a dislocation, is highly idealized. In polycrystalline alloys of
industrial importance there are many more variables that need to be considered. The
following sections will detail the most significant of these in Aluminum alloys.

2.2 Strengthening Mechanisms


A metals strength is determined by the shear stress required to activate a
dislocation. In an ideal crystal lattice this is the frictional stress. The basic principle
behind all strengthening methods is that the yield strength of a metal can be increased if
the shear stress required to activate the movement of a dislocation is increased. This is
achieved by putting obstructions in the path of dislocations. The obstructions can be
regarded as defects in a perfect crystal. Some of the more common crystal defects are
listed in Table 2-3.

Table 2-3: Various crystal defects observed in metals separated into class.
Defect class Type of defect
Self Interstitial
Interstitial Solute
Substitutional solute
Point
Vacancy
Line (1-D) Dislocations
Grain boundaries
Phase boundaries
Interface
(2-D)
Stacking faults
Precipitates Volume
(3-D)
Voids



Altering the microstructure of an alloy by adding obstructions to increase the
yield strength also commonly affects the strain hardening rate (SHR) of the material.

16
Different strengthening mechanisms will produce different microstructures which will
evolve differently when plastically deformed. If the obstructions within the lattice
structure continue to inhibit dislocation flow after initial yielding, a higher SHR will be
observed than if subsequent dislocations experience less resistance. The concept of SHR
(d/d) will be discussed in following sections.

2.2.1 Strain Hardening


The discussion up to this point has been limited to the onset of yielding in a
crystal lattice where the only resistance to dislocation motion is the frictional stress. If
this were indeed the case it would be expected that metals would exhibit ideal
elastoplastic behavior. That is, upon yielding the stress required for a dislocation to move
through the lattice would be constant. It was shown earlier in Figure 2-1 that this is not
the case and elastic-plastic behavior with parabolic work hardening is observed. In this
section the origin of this increase in flow stress with increased strain will be discussed.
Metal components are often formed or shaped at room temperature by mechanical
processes (rolling, forging, extrusion, straightening, stamping, bending, etc.).
Historically, cold shaping has been widely exploited. In order to produce a net shape
change without cracking or fragmentation, the crystal lattice needs to compensate for this
bulk geometrical change. If you recall, a dislocation produces a displacement in the
crystal lattice. The macroscopic shape change of the component is the summation of
many microscopic displacements resulting from dislocation production and flow. The
number of dislocations present in a microstructure is quantified by dislocation line length
per unit volume and termed the dislocation density. The units of dislocation density

17
(length/length
3
) can be simplified to the number of dislocations per unit area. During
deformation processing, the dislocation density () of an annealed material can increase
from a typical value of 10
10
dislocations/m
2
to over 10
16
dislocations/m
2
, which consists
of both geometrically necessary and statistical dislocations. This dramatic increase in
dislocation density affects the resistance of dislocation motion. The resistance of
dislocation motion as a result of an increased dislocation density and dislocation-
dislocation interaction is strain hardening.
There have been many models put forth to explain this increase in flow stress with
increased strain (and dislocation density). The earliest such model was proposed by
Taylor [10] from which the vast majority of subsequent models have been developed.
Although none of the proposed models are entirely accurate, Taylors captures the basic
concepts. It is assumed that a dislocation resides in a lattice with other dislocations
present. A stress is applied and the dislocation glides a distance (l) before being halted
producing a strain of:
k bl = (2.7)
where k is an orientation dependent factor. It is assumed at the dislocation distribution is
uniform and that the distance between dislocations (L) is:

1
L
= (2.8)
If only repulsion stresses are considered when dislocations interact and the minimum
distance between dislocations is taken to be half the average separation distance (L/2) a
dislocation must overcome a stress of:

1
2
KGb
KGb
L
t = = (2.9)

18
to glide, where K is a constant, which when combined with (2.7) gives the shear stress as
a function of dislocation density. Examination of equation (2.9) reveals that the flow
stress scales with dislocation density. Over subsequent years, the constant K was changed
to (which is ~0.2 for fcc and ~0.4 for bcc [9]) and the frictional stress was added to
produce the currently accepted relationship for activating a dislocation in a strained
metal, given by:

1
2
o
Gb t t o = + (2.10)
Considering all the approximations made to arrive at equation (2.10), it accounts
for the strengthening due to the increase in dislocation density resulting from strain
remarkably well, as shown in Figure 2-5 [16]. Note that the dislocation density at the
lowest resolved shear stress values match those of typical fully annealed single crystals.


Figure 2-5: Resolved shear stress plotted against dislocation density for single and
polycrystalline Copper [16].



19
The stress-strain response of single crystals was examined by Kuhlmann-
Wilsdork [17-19] to better understand the strain hardening process. A typical shear stress-
strain curve for a single crystal is shown in Figure 2-6.



Figure 2-6: Schematic representation of a shear stress-strain curve for a single crystal
[20].
Exhaustion
Hardening

CRSS


Shear Strain ()
Linear
Hardening Easy Glide

I


Stage II Stage III

III

Dynamic
Recovery

II


Stage I



The shear stress-strain curve in Figure 2-6 can be separated into four distinct regions:
- An elastic region that follows Hookes law until the critical resolved shear
strength (
CRSS
) is reached on one of the potential slip systems. This is the
externally applied stress resolved in the direction of a slip system.
- Stage I: The yield stress of the crystal has been surpassed. At this stage there is a
low dislocation density. Dislocations can easily glide along parallel slip planes
with little resistance other than the frictional stress. The strain hardening rate
(d/d=
I
) is very low, a result of overlapping dislocation stress fields from
dislocations on parallel slip planes and the formation of dipoles. Stage I is only

20
seen in single crystals where almost all of the dislocations generated by the
imposed strain can escape the bulk crystal resulting in slip steps observed on the
crystal surface.
- Stage II: The transition between stage I and II is associated with; (1) the activation
of multiple slip systems and (2) the separation between dislocations (L)
decreasing. During Stage II almost all newly generated dislocations are trapped
within their respective crystal. Dislocations are now interacting with each other
forming tangles and three-dimensional networks. This results in a greater
resistance to flow producing a segment of linear hardening [9, 21] with a much
higher strain hardening rate:
:
3000 300
I II
d G G
d
t
III
u u

( (
~ < = >
( (

u (2.11)
The strain hardening rate observed in Stage II is the inherent hardening rate in the
absence of dynamic recover.
- Stage III: As the amount of imposed strain increases the hardening starts to be
exhausted. During this stage dislocations are still being produced at the same rate
although less are being retained in the structure due to dynamic recovery. The
onset of stage III is dependent on the stacking fault energy (SFE) of the material
and deformation temperature. The onset of Stage III occurs at lower stresses in
higher SFE material deformed at increasing temperatures.

In polycrystalline metals Stage I and II are not observed. This is due to each
crystal being constrained by its surroundings, resulting in multiple slip occurring
immediately and a rapid production of dislocations. This will be discussed in more detail

21
in subsequent sections (Hall-Petch effect). The strain hardening rate of a polycrystalline
material (d/d) can be estimated from the strain hardening rate in stage III by scaling
with the Taylor factor (M) [22]:

2
d
M
d d
d o t
c
~ (2.12)
Equation (2.12) illustrates that the strain hardening rate of a polycrystalline material will
be significantly larger than a single unconstrained crystal. In fcc metals, the average
Taylor factor is approximately 3.1 which translates into a hardening rate of 9.5 times
larger in a polycrystalline structure compared to a single crystal. Stage III is strongly
effected by stacking fault energy (SFE). The higher the SFE of an alloy, the earlier the
onset of stage III. The early onset of stage III is due to cross slip being activated at lower
stresses and the easier rearrangement of dislocations to reach a stable cell size faster. A
high SFE promotes dynamic recovery and decreases the strain hardening rate in Stage III.

2.2.2 Grain Boundary Strengthening (Hall-Petch Effect)

In the previous section the emphasis was on dislocations moving through the
lattice of a single crystal under an externally applied load. While this is of great value in
understanding the complexities of how dislocations interact during stain hardening, single
crystals are of limited industrial use (with the notable exception of Ni based high
temperature turbine blades). All alloys that are used in structural or load bearing
applications are polycrystalline. That is, they are made up of an aggregate of crystals
(grains) of various orientations separated by interfaces known as grain boundaries, as

22
shown in Figure 2-7. The orientation of each crystal relative to the specimen reference is
denoted by g.


Figure 2-7: EBSD image of polycrystalline Al showing grains of various orientations
distinguished by color and separated by interfaces (grain boundaries).


Polycrystalline structures behave substantially different than single crystals when
deformed. The primary reason for this is that each grain has to cooperatively deform with
its neighbors. If it is assumed that each grain in a polycrystalline structure is free to
deform as if it were an unconstrained single crystal, voids would form in some regions
while multiple grains would occupy the same volume in other regions, as shown in
Figure 2-8 a&b. This is because an unconstrained grain will deform by activating
dislocations on only the slip system with the highest critical resolved shear stress. This
manner of deformation follows Schmids law [23] (Schmid factor =coscos, where is
the angle between the loading axis and the slip plane normal and is the angle between
loading axis and slip direction) and is directly dependent on the grains orientation. Since
industrially used metals do not immediately fracture upon yielding, and the three-
dimensional network of grains deform together, slip must occur on multiple systems at all
stages of plastic flow. It was shown by VonMises [24] that, in order for a crystal to
1
g
5
g
4
g
3
g
2
g

23
undergo an arbitrary shape change, a minimum of five independent slip systems need to
be available. The activation of these additional slip systems prevents the formation of
voids and preserves the conservation of volume. This strain incompatibility between
grains, resulting from the crystals anisotropic nature, has the additional effect of creating
a volume of increased dislocation density near the grain boundaries, as shown in Figure
2-8 c&d. Experiments have shown that in coarse grained Al the strains are higher near
the grain boundaries than in the center of the grains after deformation [26].



(d) (c)
(b)
(a)
Figure 2-8: Schematic representation of (a) a polycrystalline structure being pulled in
tension, (b) the deformation that would occur if each grain in the aggregate was
unconstrained by its neighbors and the subsequent voids and overlaps that would result,
(c) how allowing for multiple slip would eliminate the voids and overlaps (d) resulting in
a higher strain (dislocation density) near the grain boundaries [25].



It has been known for many years that the average grain size (or spacing between
grain boundaries) is a critical parameter in determining yield strength. The two primary

24
reasons for this are: (1) the increased dislocation density resulting from multiple slip as
discussed above and (2) that the mean free path of a dislocation scales with the diameter
of the grain (d) since slip systems between grains do not regularly coincide. Both of these
factors work to inhibit the flow of dislocations. The first attempt to quantify the inverse
relationship between grain size and yield strength was by Hall [27] and Petch [28]. This
was through empirical studies in which they both independently observed the inverse
square root relationship between grain size and yield strength. While the fundamental
concept of their model was correct, that grain boundaries act as barriers to dislocation
flow, many of the assumptions made by others that followed their work were not. The
two most notable inaccuracies were that dislocations are able to travel between grains and
the origin of the dislocations themselves. Although some physical phenomena in the
model are unrealistic, it served as the building block for most subsequent theories. The
basic mathematical relationship put forth by Hall and Petch still holds true. For this
reason the model will be reviewed.
It was assumed that when a strain is applied, a dislocation (Frank-Read [29])
source somewhere in the interior of a grain will start producing dislocations. After some
degree of travel (depending on the grain size), the dislocations will encounter a grain
boundary and begin to pile-up. The number of dislocations that can accumulate between a
grain boundary and a dislocation source was estimated by Eshelby [30] to be:

4
S
d
n
Gb
ott
= (2.13)
where
s
is the average resolved shear stress in the plane and 1. In order for the
dislocation at the tip of the pile-up to pass through the grain boundary and into the
adjacent grain it must surpass a critical stress given by:

25

2
4
s
c s
d
n
Gb
ott
t t = = (2.14)
Recalling that the resolved shear stress is the applied stress without the frictional stress:

s o
t t t = (2.15)
allows equation 2.14 to be further simplified and rearranged to produce:

1
1
2
2
4
c
o o
Gb
k d
d
o t
t t t
t
y

| |
= + = +
|
\ .
(2.16)
which is know as the Hall-Petch relationship, with all the constants grouped into the term
k .
y
Graphically, what this relationship shows is that yield strength will vary linearly
(with a slope of k
y
, referred to as the locking parameter) with the reciprocal square-root
of the grain size. This relationship holds true for a large number of alloy systems, as
shown in Figure 2-9 [31]. Examination of Figure 2-9 reveals that strengthening by grain
refinement is much more effective in some alloy systems than others. The slope of the
aluminum data (shown in dark green) is significantly less then other alloys, as shown by
the values given in Table 2-4 [31], resulting in less of a strengthening benefit from
reducing the grain size. It should be recognized that the addition of 3.5%Mg increases
both Hall-Petch parameters. It has been postulated that this increase in Hall-Petch
parameters is a result of a decrease in the SFE which helps to inhibit dynamic recovery.
The data presented in Chapter 4 of this document will more strongly support that it is the
solute-dislocation interaction itself inhibiting dynamic recovery and not related to SFE.



26
0
100
200
300
400
500
600
0 5 10 15 20 25 30
d^-1/2 (mm^-1/2)
S
t
r
e
s
s

(
M
P
a
)
1. 304 S.S. Annealed (1% Offset) 2. 304 S.S. Rolled (1% Offset)
3. Brass (Y.S.) 4. Copper (Y.S.)
5. "Ferrovac E" Steel (Y.S.) 6. 0.05 C Steel (Y.S.)
7. 0.09 C Steel (Y.S.) 8. 0.13 C Steel (Y.S.)
9. 0.15 C Steel (Y.S.) 10. 0.20 C Steel (Y.S.)
11. Nickel (Y.S.) 12. Molybdenum (Y.S.)
13. Aluminum (Y.S.)

Figure 2-9: Hall-Petch plot for various alloys showing a linear relationship between yield
able 2-4: Hall-Petch parameters for various alloys [32].
stress and the reciprocal square of grain size, with a slope of k
y
[31].



T

y
k
y

y
k
y

Material Material
(MPa) (MN/m^1.5) (MPa) (MN/m^1.5)
bcc alloys fcc alloys
Mild Steel 6 0.74 Copper, e=0.005 .5 0.11 70. 25
Mild Steel, e-0.1 29 11 4.18 0.39 Cu-3.2%Sn 1.79 0.19
Swedish Iron 47.07 0.71 Cu-30%Zn 45.11 0.31
Fe-18%Ni,
e=0.002 650.14 0.22 15.69 0.07
Aluminum,
e=0.005
Chromium g 178.47 0.90 Al-3.5%M 49.03 0.26
Molybdenum 05 107.87 1.77 Silver, e=0.0 37.26 0.07
Tungsten 640.33 0.79 Silver, e=0.002 23.53 0.17
Vanadium 318.7 0.30 Silver, e=0.20 150.03 0.16
Niobium 68.64 0.04
hcp alloys
Low-carbon steel 0 Magnesium 6.86 0.28 ---- .307
Armco iron ---- 0.583 Titanium 78.45 0.4
Zirconium, e=0.002 0 29.42 .25
Beryllium 21.57 0.41
Zinc ---- 0.22


27
The Hall-Petch relationship predicts that the lower bound of the Hall-Petch plot
(the y-intercept) will be the frictional stress of the alloy and the upper bound will
approach extremely high stress values on the order of the theoretical strength.
Experiments [33-38] have shown that there is a critical grain size (d<0.5) at which this
relationship breaks down, as seen in Figure 2-10.



Figure 2-10: Hall-Petch plot showing a change in slope (k
y
) at small grain sizes. Note
that the vertical axes is Vickers hardness rather than yield stress [35].


Another reason the Hall-Petch relationship breaks down at smaller grain sizes is
that the physical metallurgy used to rationalize the strengthening mechanism is incorrect
or at best incomplete. This is true, not only for the original model, but for some of the
models that followed as well [39].
A more recent model proposed by Meyers and Ashworth [40] appears to not only
explain the strength-grain size relationship over the typical range explored but, also
accounts for the breakdown at smaller grain sizes. Building on previously established
information, Meyers and Ashworth considered a grain in an anisotropic aggregate which,

28
when deformed, would cooperate with the surrounding grains; therefore, activating
multiple slip systems. As a result of the anisotropic nature of metals, an incompatibility
stress arises between adjacent grains. It is estimated that the interfacial shear stresses
resulting from this incompatibility are on the order of three times greater than the
resolved shear stress within the grain. This increased stress at the grain boundaries results
in microscopic yielding in which the dislocation sources are the grain boundaries (rather
than within the interior of the grains). As dislocations are emitted from the grain
boundary, they travel a short distance into the volume of the grain until the short range
stress pushing them diminishes. When subsequent dislocations are emitted they approach
the previously emitted dislocations and are repelled by them. This limits the thickness of
the strained layer, as shown in Figure 2-11. A steady state is reached as the dislocations
pile-up and the stress builds, promoting cross-slip in this region of high dislocation
density.


Figure 2-11: Schematic representation of (a) an annealed polycrystalline structure and (b)
the resulting local plastic flow in the grain boundary region (microyielding) at the start of
deformation followed by (c) the formation of a hardened region with an increased
dislocation density that effectively reinforces the microstructure [40].

29
As the applied stress continues to increase, the hardened region around the grain
boundaries will initially carry the stress. This will continue until large scale yielding
occurs, at which point other dislocation sources become activated and the interior of the
grains deform.
The Meyers-Ashworth model not only appears to more accurately represent the
evolution of the microstructure during yielding, but also accounts for the breakdown of
the Hall-Petch effect at smaller grain sizes. As the grain size decreases, the grain
boundary surface area per unit volume increases resulting in greater resistance to
dislocation flow. More importantly, from the point of view of the Meyers-Ashworth
model, as the grain boundary surface area increases so does the volume of the hardened
region around the boundaries. Eventually, the grain size reaches a point where this
hardened region is a substantial volume of the grain. Once this point is reached, further
reducing the grain size will have a diminished strengthening effect. This seems to be a
plausible explanation of the Hall-Petch relationship and the lower k
y
values observed at
small grain sizes.

2.2.3 Solution Hardening

High purity metals are rarely used when enhanced mechanical properties are
required. The majority of commercially used metals are alloys which are composed of a
base metal that makes up the matrix (solvent) and an alloying element or elements
(solute) that are added for improved properties. These elements can be added in
combinations and ratios that promote the formation of additional phases, which will be
discussed in the next section, or remain in solid solution. The solute atoms that remain in

30
solid solution inevitably have a different atomic radius then the base atoms. The
mismatch in atomic size determines where in the lattice the solute atoms will reside. In
general, if the sizes of the solute atoms are on the same order or larger than the atoms of
the base element, the solute atom will reside at a lattice site normally occupied by a base
atom, forming a substitutional solid solution. If the solute atoms are much smaller than
the atoms of the base element, the solute atom will usually reside in-between lattice sites
forming an interstitial solid solution.
The mismatch in size between the solute atoms and base atoms causes the lattice
to distort in order to accommodate the volumetric change. This distortion produces a
short range elastic stress field. Substitutional solute atoms produce either a positive or
negative symmetric-hydrostatic stress depending on their relative size, as shown in
Figure 2-12.


(a) (b)
Figure 2-12: Schematic representation of a solute atom (a) producing a negative
(compressive) stress field, and (b) a positive (tensile) stress field [41].


The stress field produced by the solute atoms interacts with the dislocation stress fields
and impedes their flow. The magnitude of this impediment to dislocation flow scales with
the magnitude of misfit () which is a function of the lattice parameter and solute
concentration. In order for this type of parelastic interaction to occur, the solute atom
does not have to lie directly in the slip plane. The stress field of the dislocation is long-

31
range and will interact with the short-range stress field of a solute atom if it lies above or
below the slip plane. If the solute atom resides above the slip plane, where the dislocation
stress field is negative, there are two possible scenarios. Either the solute atom will: (1)
have a negative stress field associated with it and the two fields will repel each other
(requiring an increased shear stress to overcome the obstacle) or inversely, (2) the solute
atom will have a positive stress field and they will attract forming a Cottrell
atmosphere [43], effectively pinning each other. If the solute resides below the slip
plane the opposite would occur, as shown in Figure 2-13.


(a)
(c) (b)
Figure 2-13: Schematic representation of (a) an edge dislocation exerting a negative
stress field above the slip plane and positive below, (b) smaller substitutional solute
atoms with a positive stress field interacting with an edge dislocation to mutually relax
their fields, and (c) larger substitutional solute atoms with a negative stress field
interacting with an edge dislocation to mutually relax their fields [41].


Unlike an edge dislocation, whose stress field has both hydrostatic and shear
components, the stress field of a screw dislocation is pure shear. A screw dislocation does
not have a volumetric change associated with it, although the atomic bonds are distorted
in the region. As a result the size effect discussed in the previous paragraph does not
inhibit the flow of a screw dislocation. This is because there is not a symmetric portion of
the dislocation stress field to be relaxed by the symmetric stress field of the substitutional

32
solute atoms. The stress fields of screw dislocations interact with non-symmetric stress
fields which are most often associated with interstitial atoms. As a point of clarity, solute-
dislocation interaction depends on whether the stress field associated with the solute atom
is symmetric or non-symmetric, not necessarily if the atom is interstitial or substitutional.
Ideally, the structure will contain atoms that effectively interact with both edge and screw
dislocations.
The method by which solute atoms impede dislocation motion that has thus far
been discussed is elastic field interaction. There are two other mechanisms by which this
occurs, modulus interaction and stacking-fault interaction. It is unlikely that a solute atom
and a matrix atom will have the same shear modulus, as a result, the effective modulus of
the matrix around the solute atom is locally changed. This, in-turn, alters the local energy
and elastic field experienced by a dislocation. The magnitude of this dielastic interaction
() is a function of the elastic shear modules and solute concentration. Since the elastic
stress field in this situation is produced by a local change in the effective bulk modulus it
doesnt matter where the solute atom resides relative to the slip plane. The resulting
change in the elastic field will be asymmetric and affect both edge and screw
dislocations.
The combined effects of the parelastic interaction resulting from the atomic
mismatch and dielastic interaction resulting from the local change in elastic modulus was
experimentally demonstrated by [43-45]. Single crystal critical resolved shear stress
values for various binary Cu alloys were measured as a function of solute concentration.
It was shown that some alloying elements more effectively strengthened the crystals than
others, although all the binary alloys exhibited a 2/3 power dependency on the solute

33
composition, as shown in Figure 2-14. The relative effectiveness of each alloying
element can easily be determined by the slope (d
CRSS
/dc
2/3
) of each data set. The
mismatch and modules effects were described using a combined interaction parameter
proposed by Labusch [46, 47] and given by:

2 1 2 2 2
) ( q o o c + =
L
(2.17)
where is a geometric variable dependent on dislocation curvature. The positive
correlation between the magnitude of atomic mismatch and local modules disruption on
the effectiveness of solute strengthening can be seen in Figure 2-15.




Figure 2-14: Critical resolved shear stress of single crystals tested at room temperature
(plateau stress) plotted against solute concentration to the 2/3 power [48].


Solute Concentration - c
2/3

c
r
s
s

(
K
N
/
c
m
2
)


34


(
M
P
a
)

2
/
3
/
d
c
C
R
S
S
d

2 1 2 2 2
) ( q o o c + =
L
Figure 2-15: The slope of each binary alloy data set shown in Figure 2 14 plotted against
the Labusch combined interaction parameter [48]


It is also possible that the solute atoms may have a higher solubility in a different
crystal structure than the matrix. For instance, a packing sequence provided by a stacking
fault. In this situation the solute atoms will preferentially concentrate in this fault,
effectively lowering the SFE. As the SFE decreases cross slip, dislocation flow, and
dynamic recovery become more difficult.
The macroscopic mechanical response to the increase in flow stress required to
move a dislocation through a structure strengthened by solute is shown in Figure 2-16.
As the amount of solute increases so does the yield strength. It is also important to note
that the strain hardening rate increases with increasing solute. This shows that solute
strengthening is present through the entire deformation process. Solute strengthening is
most closely related to an increase in frictional stress.


35
0
100
200
300
400
500
600
0 0.5 1 1.5 2 2.5
Rolling Strain
T
e
n
s
i
l
e

Y
i
e
l
d

S
t
r
e
s
s

(
M
P
a
)
0
10
20
30
40
50
60
70
80
T
e
n
s
i
l
e

Y
i
e
l
d

S
t
r
e
s
s

(
K
s
i
)
3Cu-1Mg
2Cu-0.5Mg
1.5Cu-0.5Mg
1Cu-0.5Mg
Al 99.99%
Figure 2-16: Tensile yield strength plotted against cold rolling strain showing both an
increased yield point and strain hardening rate for alloys of increase levels of solute [5].


A quantitative description of the relationship between solute concentration and
flow stress is not agreed upon, so one will not be derived. It is estimated that the
strengthening increment (referred to as increment because the frictional stress is not in
the equations) due to a solute atom with a non-symmetric stress field is given by:

1
2
non sym
c
Gb
b
t |

| |

A ~

|
\ .
|
|
(2.18)
where c is the solute composition and is a constant ~1. For a solute atom with a
symmetric stress field the strengthening increment is given by:

3 2 1 2
700
s
sym
c
Gb
b
c
t A ~ (2.19)

36
where
s
is a measure of the change in lattice parameter. Note the absence of the
s
term
in the non-symmetric equation.
The complexity of accurately deriving expressions for the increase in flow stress
due to solute atoms arises for numerous reasons. There are multiple ways in which
dislocation-solute interaction can occur and no method of distinguishing between them.
Stress is a state function while strain is path dependent, making the dislocation structure
path dependent. Also, an often ignored complexity is that a dislocation will not take the
form of a straight line, but rather bend and twist to minimize its energy (similar to the
assumption Zener made in considering a grain boundary interacting with particles [49]
and later corrected by [50]). These are just a few of the complexities in deriving a unique
relationship between flow stress and solute concentration. The important relationship that
comes out of equations 2.18 and 2.19 is that the increased flow stress is proportional to
Gb and scales with the square root of solute composition, which is experimentally
observed, as shown in Figure 2-17 [9]. It should be noted that while the square root
relationship is most often quoted, multiple investigators have produced data sets showing
a 2/3 relationship such as the one previously shown in Figure 2-14.



37

Figure 2-17: Increased flow stress varying with the square root of solute composition for
(a) substitutional atoms and (b) interstitial atoms [9].


2.2.4 Precipitation Hardening

Logic would dictate that if solute atoms effectively increase the resistance to
dislocation flow, then if these atoms concentrate and form second phase particles this
would further impede the movement of dislocations. This is the basic principle behind
precipitation hardening, commonly exploited in Al and Ni based alloys and to a lesser
extent in steels (often referred to as secondary hardening steels or PH steel).
Precipitation is the most common phase transformation in alloys that retain the
same crystal structure at all temperatures in their solid state, as does Al (fcc). The
reaction occurs by raising the bulk metal to an elevated temperature where the alloying
elements are soluble in the matrix. The alloy is then rapidly quenched to room
temperature, locking the solute atoms in a supersaturated solid solution, as shown in
Figure 2-18.


38

Figure 2-18: Al-Cu phase diagram showing a possible solution heat treat and quench to
produce a supersaturated solid solution [41].



At this lower temperature the alloying elements are no longer soluble in the matrix and
fall out of solution forming second phase particles. These second phase particles are
precipitates. The precipitation reaction can be expressed as follows:
+
where is the metastable supersaturated solid solution, is the more stable solid solution
with the same crystal structure as , and is the precipitate which may be either stable or
metastable. The precipitates are different in chemical composition and/or crystal
structure. The actual phase that forms may not be the most energetically stable phase. The
precipitation reaction is a combination of the thermodynamic driving force and the
kinetics of the system [51]. The phase that appears first will be the one with the lowest
barrier to nucleation, which is not always the most thermodynamically favorable. This is
clearly shown in Figure 2-19 and Table 2-5; where in the Al-Cu alloy precipitation
sequence the most thermodynamically unfavorable phase forms first and the more
favorable phases follow sequentially.

39

2-19: Schematic representation of the molar free energy diag

Figure ram for the Al-Cu
system [52].

cipitation sequence for various precipitation hardening alloys [31].
ects
how they interact with dislocations. In general, the first particles to nucleate within a

Table 2-5: Pre


The sequence in which precipitates form is relevant because it indirectly aff
grain will be either coherent or semicoherent. This is because coherency between the
matrix-precipitate lattices reduces the barrier to nucleation by having relatively
(compared to an incoherent interface) lower interface energy. Likewise, precipitates that

40
first form on grain boundaries are usually more thermodynamically favorable then ones
simultaneously forming in the interior of the grain because a high energy interface
already exists lowering the barrier to nucleation. These grain boundary particles are
typically incoherent and do not contribute to strengthening.
Precipitates interact with dislocations in a similar manner as solute atoms do.
There is an elastic stress field associated with precipitates and they change the local
Figure
precip the
precip

increased. This increases the energy of the system by:
modulus of the matrix; although, these are not the primary strengthening mechanisms in
precipitation hardening. The strengthening provided by precipitates is a result of them
residing on slip planes, acting as physical barrier to dislocation flow. Consider a
dislocation approaching a spherical particle that is bisected by a slip plane, as shown in
Figure 2-20.


2-20: Schematic representation of (a) a dislocation approaching a spherical
itate that lies on its slip plane and (b&c) shearing it as it passes increasing
itate-matrix interface which increases the energy of the system [9].


If the dislocation passes through the precipitate, shearing it, the particle-matrix interface
is
2
s
U b r t ~ (2.20)

41
where
s
is the surface energy. A relationship for the force required to shear the particle
can be determined by recalling that force is the rate of change of energy with respect to
distance, producing the relationship:

max
s
r b dU
F
dx b
r s
t
t = = = (2.21)
This force is being applied over an area equivalent to the Burgers vector along the
precipitate spacing (), producing a flow stress of:
,
s s
Sh sh
r f r
f
b b
t t
t t

= = =

(2.22)
where f is the volume fraction. Examination of equation 2.22 reveals that the flow stress
should continually increase as a precipitate coarsens (Ostwald ripening). As illustrated in
the aging curves shown in Figure 2-21, it is known that this is not the case.



Figure 2-21: Characteristic aging curves for an Al-Cu alloy [41].




42
During the aging process, an alloys strength increases from an under-aged
condition to a peak aged condition, where the strength is the highest, and finally to an
overaged condition, where the strength decreases once again. As time at temperature
increases, the volume fraction of precipitates increases until it reaches a constant value
which depends on the alloy composition and phase diagram. At this point Ostwald
ripening occurs [53], increasing the precipitate radius and spacing. Simultaneously, the
precipitates within the grains, which are the only ones that contribute to the
strengthening, are continuing the precipitation sequence (from typically semi-coherent to
non-coherent crystal structures) to lower the energy of the system. In general, the more
stable the phase, the less coherent it is. All of these factors contribute to the shape of the
aging curve.
As the microstructure evolves during the aging process, the precipitates are either
no longer able to be sheared due to incompatible Burgers vectors or the stress to shear the
precipitates is too high. Once the precipitates become totally incoherent, the interface
becomes similar to that of a grain boundary and dislocations are no longer able to travel
through them. Also, as Ostwald ripening progresses, the radii of the precipitates become
too large for the flow stress to shear. Instead, the dislocations by-pass the precipitates by
a mechanism know as Orowan bowing [54], as show in Figure 2-22.


43

Figure 2-22: Schematic representation of Orowan bowing showing (a) a dislocation
approaching an array of precipitates, (b&c) bowing around them, and (d) by-passing them
leaving a dislocation loop.


As in the shearing mechanism, the dislocation approaches the precipitate lying in
its slip plane, but instead of progressing through its lattice it starts to bow past it. As the
flow stress increases, the bowing becomes more extreme until a critical point is reached.
The Burgers vectors on the backside of the particle become increasingly aligned with
each other, although with opposite signs and cancel out. A dislocation loop is left around
the non-shearable precipitate as the gliding dislocation by-passes it. Similar to how a
dislocation bows and interacts within a Frank-Read source [29] the stress (
Ow
) required
to bow and by-pass a precipitate is given by:
,
Ow Ow
Gb r Gbf
f r
t t

= = = (2.23)
The shear stress necessary for Orowan bowing has an inverse relationship to r as the
shearing mechanism. As a result, it requires less stress for a dislocation to shear a small
particle and by-pass a large one, as shown in Figure 2-23.


44

Figure 2-23: Shear stress plotted against precipitate radius for both the shearing process
and the by-pass process, showing a switching between mechanisms at some critical
radius [9].
{
{
}
}
}
{
}
{
}
{



Switching from shearing to by-pass explains the overaging response observed and
the shape of the aging curves. When the precipitation sequence begins out of a
supersaturated solid solution, the radii of the second phase particles are initially small. At
this stage, the shearing mechanism will dominate because it requires the least amount of
stress. As time goes on, the mean particle radius increases until a critical point is reached
(r
c1
, r
c2
, r
c3
) where it requires less stress to by-pass a particle and Orowan bowing
becomes the primary strengthening mechanism. The effect of volume fraction (f
1
< f
2
)
and modulus (B>A) is also shown in Figure 2-23. The peak-aged condition occurs either
at this crossover point or right before it. The highest possible yield strength is achieved
when both mechanisms are active.
The method by which dislocations interact with precipitate particles also effects
the SHR. It was shown by Doherty et al. [5] that heat treatable Al alloys strain harden
more rapidly when the solute remains in solution (naturally aged condition) than when it

45
is peak-aged. At extended aging times, when the strengthening method has switched from
shearing to by-pass, this trend is reversed for the initial ~2% strain, as shown in Figure
2-24.



(a)

(b)
(c)
Figure 2-24: (a) Critical shear stress of Al-1.6%Cu specimen tested at room temperature
after being aged for various times at 190C. (b) Room-temperature stress-strain curves
after aging for select times at 190C. (b) SHR (at 2% strain) plotted against aging time
[55].


Peak-aging is reached in this alloy after ~40 hours at 190C. At times before peak-aging
the yield stress is increasing and the SHR is decreasing. After ~30 hours the dislocations

46
are starting to by-pass precipitates resulting in a continuously decreasing yield strength
and increasing SHR. It should be noted that the increase in SHR observed in the overaged
material spans a relatively small strain interval (2%). It will be shown in the experimental
results contained within Chapters 5 of this document that the SHR in the overaged
condition is exhausted more rapidly than in the underaged or peak aged condition. This
more rapid exhaustion of the SHR with increasing aging has also been reported by [143].
At the larger strains being investigated in the current study, the SHR characteristics of the
overaged material are less desirable. The increased SHR in the overaged condition is
partially due to the dislocation loops left behind after a dislocation passes, which are not
present when shearing occurs, as shown in Figure 2-25.


(b) (a)

Figure 2-25: (a) Spherical ' precipitates in a Ni-19%Cr-6%Al alloy aged at 750C for
540h by dislocation flow being sheared [56] and (b) dislocations bowing between
precipitates leaving loops behind in an Al-0.2%Au alloy deformed 5% after solutionizing
and aging at 200C for 60h. [57].



This produces a stress field around the precipitates that repels approaching dislocations
and effectively reduces the particle spacing. Both of these phenomena will increase the
SHR by further inhibiting subsequent dislocations to flow.


47
2.2.5 Total Strengthening Effect

The effect of various strengthening methods has been reviewed as if they are
independent of each other. This has produced a set of relationships that allow for the
estimation of each, summarized in Table 2-6.


Table 2-6: Summary of strengthening mechanisms.
Strengthening
Method
Interaction Description Eq. #
Frictional Stress
(
o
)
Dislocation-Lattice
2
2
(1 )
a
w
b
b
o
Ge Ge
t
t
u
t

~ ~
2.6
Strain
Hardening
(
S-H
)
Dislocation-Dislocation
1
2
o
Gb t t o = +
2.10
Boundary
Strengthening
(
BS
)
Dislocation-Crystal
interface
1
2
o y
k d t t

= +
2.16
Solution
Strengthening
(
SS
)
Dislocation-Solute
atom
o
Gb c t t +
2.18
&
2.19
Dislocation-Precipitate
(By-pass)
,
Ow Ow
Gb r Gbf
f r
t t

= = =
2.23
Precipitation
Hardening
(
Ow
&
sh
)
Dislocation-Precipitate
(Shearing)
,
s s
Sh sh
r f r
f
b b
t t
t t

= = =

2.22


In realistic microstructures, multiple strengthening methods obstruct dislocation
flow. Take for example a precipitation hardened Al alloy that is subjected to a shaping
process. As load is applied and metal starts to flow, dislocations are not only interacting
with the strengthening precipitates, but also any solute that remains in solution, grain
boundaries, and other dislocations. In all of the shear stress relationships the frictional
stress is either simply added on or the strengthening is given in an incremental form, as it
is for precipitation hardening. The individual strengthening relationships given in Table

48
2-6 do not provide any insight into how these complex mechanisms interact with each
other or how the resulting flow stress is affected. It has been empirically shown that the
frictional stress and boundary stress are additive while the other mechanisms add as the
sum of the squares, as shown below:

( )
1 1
2 2
2 2
(
Total o y S H SS Ow sh
k d t t t t t t

= + + A + A + A + A
2
) (2.24)
It should be noted that, while equation 2.23 may be the best estimate of total
combined shear strength, it is only an estimate. Strengthening is a dynamic path
dependent process that spans multiple length scales with multiple strengthening
mechanisms that can interact with each other to both reinforce and detract from the
overall resistance to dislocation flow.

2.3 Material Failure by Fracture

The previous sections dealing with yield strength and plastic deformation may
lead one to conclude that a metallic component is safe from failure provided that the
service stresses are below the yield point. A proper design methodology for a load
bearing component must consider more than just plastic yielding as its failure criteria.
There are numerous examples of structures that have failed due to fracture, often with
tragic consequences. Fracture is a particularly dangerous failure mode because it often
occurs catastrophically, without warning, and at stress levels safely below the yield point
of the material.
In order for a solid to fracture or fragment into multiple parts, the cohesive stress
(
c
) that holds the atoms together must be surpassed. The cohesive stress of a crystalline

49
material can be calculated in a similar manner as the theoretical shear stress was for
yielding. Assuming a perfect crystalline structure, atoms will arrange themselves at
equilibrium spacings (x
o
) that minimize the potential energy of the system. The
equilibrium spacing coincides with a state of zero applied force, as shown in Figure 2-26.



2

Figure 2-26: Potential energy and applied force plotted against inter atomic spacing for
two adjacent atoms [58].



If the shape of the force-displacement curve is approximated as a Sine function and the
small angle approximation is employed it can be described by:
2 2
sin
c c
x x
F F F
t t

| | | |
= =
|
\ . \ .

|
(2.24)

50
where the cohesive force is given by F
c
, the distance from equilibrium is x, and the wave
length is . This equation can then be written in terms of stress by multiplying each side
by the number of bonds per unit area, producing the relationship:

2
c
x t
o o

|
=

\ .
|
|
(2.25)
If it is assumed that the imposed strain is elastic then the following relationship can be
employed:

o
x
E E
x
o c
| |
= =

\ .
|
(2.26)
Combining and simplifying equations 2.25 and 2.26 to solve for the cohesion stress
gives:

2
,
2
c
o o
c
x x
E
E
x x
t
o o
t
| |
| |
= =
| |
\ .
\ .
(2.27)
which can be further simplified to:
,
2
o c
E
x

o
t
~ = (2.28)
When fracture occurs, atoms that where once in contact with similar atoms are
now in contact with the environment. The work (area under the stress-displacement
curve) required to create this free surface can be estimated as:

2
0
2
sin
c c
x
Work of fracture dx

t
s
o o
t
= = =
}
(2.29)
and is equal to the increase in surface energy (
s
) of the system. Recognizing that the
formation of a crack creates two surfaces and combining equations 2.28 and 2.29 predicts
the cohesive strength of a brittle crystal to be:

51

s
c
o
E
x

o = (2.30)
The cohesive strengths predicted by the relationships derived above overestimates
those experimentally observed by three to four orders of magnitude. Among the many
reasons for this over estimation, the two most notable ones are: the assumptions of a
perfect crystal and a completely brittle material. It was revealed during the discussion of
strengthening mechanisms and yielding that there are always defects present in crystals.
Depending on the type of crystal (or mechanical/geometric) defect it may behave as a
stress raiser, magnifying the local stress levels and causing them to surpass the cohesive
stress. The assumption of a completely brittle material is also unrealistic. The work done
during fracture is comprised of much more than just that required to create free surfaces.
In metals there is always some degree of plastic work at the crack tip as well as the
emission of sound, heat, and possibly even light. All of these physical phenomena require
the expenditure of work.

2.4 Fracture Toughness

A materials toughness is its inherent resistance to fracture. The field of fracture
mechanics is focused on understanding, quantifying, and predicting the failure of
materials. Through the use of various toughness parameters, a designer is able to
determine weather a component will fail under a given load configuration assuming a
maximum flaw size. This fracture sensitive approach design methodology is depicted in
Figure 2-27. These toughness parameters become inherent material properties when their
critical values are determined in a manner that is geometry insensitive.

52
Fracture toughness parameters can be separated into two distinct groups:
parameters that quantify energy to fracture and parameters that determine the magnitude
of stress at fracture. There have been several efforts aimed at developing laboratory
procedures that obtain valid toughness parameters from a material, some of which are
given in Table 2-7. The following sections will address some of the most relevant
toughness parameters.



Figure 2-27: Fracture sensitive approach design methodology depicting the
interdependence of toughness, applied stress, and flaw size. The toughness parameters are
a result of the microstructure of the material.
Applied stress
(
app
)
Flow size
(a)
Toughness
Parameters
(G, K, J, R
J
)


















53
Table 2-7: Subset of standard test methods available to extract toughness parameters [59].
Standard Title
ASTM E 399 Linear-Elastic Plane-Strain Fracture Toughness K Ic of Metallic Materials
ASTM E 561 K-R Curve Determination
ASTM E 813 JIc, A Measure of Fracture Toughness
ASTM E 1152 Determining-J-R-Curves
ASTM E 1290 Crack-Tip Opening Displacement (CTOD) Fracture Toughness Measurement
ASTM E 1304 Plane-Strain (Chevron-Notch) Fracture Toughness of Metallic Materials
ASTM E 1737 J-Integral Characterization of Fracture Toughness
ASTM E 1820 Measurement of Fracture Toughness
ASTM E 1823 Standard Terminology Relating to Fatigue and Fracture Testing
ASTM B 645 Linear-Elastic Plane-Strain Fracture Toughness Testing of Aluminum Alloys
ASTM B 646 Fracture Toughness Testing of Aluminum Alloys
ISO 12737 Metallic materials-Determination of plane-strain fracture toughness


2.4.1 Griffith Energy Balance and Energy Release Rate

The first attempt to reconcile the discrepancies between the theoretical cohesive
strength and the observed fracture strengths was made by A. A. Griffith [60]. Griffith
employed the first law of thermodynamics (conservation of energy) to describe the
conditions for crack growth in a crystalline body. Throughout this text, crack growth and
crack extension refer to the growth of a fracture crack, not the growth of a fatigue crack.
The Griffith energy balance states thatError! Bookmark not defined. the total change in
energy (dE) of the system must be equal to the sum of the change in potential energy
(d) and work required (dW ) to create a free surface for an increment of fractured
surface area (dA) given by the following expression:
S
0 = +
H
=
dA
dW
dA
d
dA
dE
S
(2.31)
The first law of thermodynamics states that energy can not be created or
destroyed, it can only change forms. Given this law and defining the system as the

54
loaded body allows for the net change in total energy to be set equal to zero which
simplifies equation 2.31 to produce:

dA
dW
dA
d
S
=
H
(2.32)
What this relationship simply states is that the increase in surface energy that results from
an incremental increase in fracture surface is balanced by the reduction in potential
energy (elastic strain/external forces) that results from the same incremental increase in
fracture surface.
Griffith applied the energy balance principle to an infinitely wide plate containing
a center crack, as shown in Figure 2-28. With the use of a stress analysis derived by
Inglis [61], it was shown that the potential energy could be described by:

E
B a
O
2 2
to
H = H (2.33)
where
o
is the potential energy of the uncracked plate, a is the crack length, B is the
thickness of the plate, and E is the elastic modulus.


55

Figure 2-28: Schematic representation of an infinitely wide plate containing a center
crack subjected to a constant stress.



Differentiating equation 2.33 with respect to the fractured surface area produces:

E
a
dA
d
2
to
=
H
(2.34)
The work required to create the crack is the product of the fractured surface area (A) and
surface energy given by:

S S s
A aB W 2 ) 4 ( = = (2.35)
Differentiating equation 2.35 with respect to the fractured surface area produces:

S
S
dA
dW
2 = (2.36)
Combining equations 2.34 and 2.36 provides the relationship:
B
o
A
2a
o
2b

56

a
E
S
y
t

o
2
= (2.37)
which is a more useful relationship for design.
The Griffith criterion for crack extension was derived assuming a totally elastic
material with a sharp crack. Both of these assumptions are problematic when dealing with
metals. It is known that even in alloys that exhibit brittle behavior there is always some
degree of plastic deformation. This plastic work not only increases the materials
resistance to crack extension but also blunts the crack tip. As a result of these
assumptions, the Griffith criteria as described thus far, is a necessary condition for
fracture but may not be sufficient for fracture to occur.
In an attempt to account for the dislocation activity during crack extension, the
expression derived by Griffith was modified to include plastic work (
p
) [62, 63] to
produce:

S p
p S
y
a
E

t

o >>
+
= ,
) ( 2
(2.38)
with the plastic deformation energy being much greater than the surface energy. This
expression was later generalized to account for all work necessary for fracture (W
f
) given
by:

a
EW
a
E
f wf
y
t t

o
2 2
= =

(2.39)
Although all of the work terms are now summed into a single generic term each
individual term is difficult to account for and/or measure. Because of this practical
limitation, the expression is of limited use in the prediction of fracture.

57
When the analysis above is reframed in terms of the energy available for crack
extension it takes on a more practical form. The energy release rate or crack driving
force, defined by Irwin[64], is given by:

dA
d
G
H
= (2.40)
As previously described (refer to equation 2.34), the crack driving force for a wide plate
subjected to constant stress with a center crack is given by:

E
a
G
2
to
= (2.41)
and crack extension (the creation of new fracture surface) will occur at a critical point
where:
f
f
c
W
dA
dW
G 2 = = (2.42)
This critical value is a measure of the material's resistance to crack extension and a
measure of fracture toughness. Combining equations 2.39 and 2.42 produces the
relationship:

a
EG
y
t
o
2
= (2.43)
Note the similarity between equations 2.39 and 2.43. The significant difference between
the relationships is that W
f
is a collection of terms measuring energy dissipation while G
is a single term representing the energy source, which is more easily measured. Stated
slightly differently; G (the crack driving force) is the energy being applied to the material
and available for crack extension, W
f
is a complex term describing the work required to
extend a crack (a materials resistance to crack extension). Crack extension will occur
when:

58
f C
W G G 2 = = (2.44)
Crack extension can be either stable or unstable. It should also be pointed out that
2W
f
has been treated as a material constant throughout this discussion. In many metals it
is not. These two topics will be discussed in later sections when resistance curves are
introduced.

2.4.2 Stress Intensity Approach

The field of fracture mechanics developed out of the early work of Griffith and its
modifications. Arguably, the two most significant contributions that came out of
Griffiths work was the assumption that all materials inherently contain flaws that act as
local stress raisers and that a materials resistance to fracture can be quantified. Linear
elastic fracture mechanics (LEFM) approaches the problem of fracture in a different
manner than Griffiths energy approach. LEFM describes the magnitude of the local
stress field at the crack tip and uses a stress intensity factor (K) to quantify it.
The fundamental principle of fracture is that a stress raiser (tip of a crack)
magnifies an applied stress to exceed the cohesive strength of the material, causing it to
fracture. Inglis [61] described the effect of an elliptical hole in an infinitely large flat
plate, as shown in Figure 2-28. It was shown that the stress at point A is given by:

|
.
|

\
|
+ =
b
a
A
2 1 o o (2.45)
This equation is more conveniently expressed in terms of the radius of curvature ():

|
|
.
|

\
|
+ = =

o o
a
a
b
A
2 1 ,
2
(2.46)

59
which can be simplified due to the relative length dimensions:

o o
a
a
A
2 , = << (2.47)
The rearrangement of equation 2.47 produces the stress concentration factor (k
t
):
= =
A
A
t
a
k o
o
o
, 0 , 2 (2.48)
Examination of this relationship reveals that as the crack tip becomes sharper the stress
approaches infinity. This seemingly unrealistic result, that materials can withstand
infinite stresses, is what motivated Griffith to approach the problem using an energy
balance. In all materials the smallest possible crack tip radius is an atomic radius and
while very small, its not infinitely small. In metallic systems there is also almost always
some degree of crack blunting under stress. Stress concentration factors have been
derived and tabulated for various geometries [65, 66].
The stress concentration factor describes the effect of crack geometry on the local
crack tip stress level. More importantly, it reveals the inverse relationship between stress
levels and the sharpness of a flaw. This term does not carry any information regarding the
stress state. A crack tip experiences biaxial and triaxial stresses when subjected to a far-
field stress (this will be discussed in greater detail in the following section). The stress
fields ahead of a crack tip in a linear elastic isotropic material were described by [67, 68]
as:

(

|
.
|

\
|

|
.
|

\
|
=
2
3
sin
2
sin 1
2
cos
2
2 1
u u u
o o
r
a
xx
(2.49a)

(

|
.
|

\
|
+
|
.
|

\
|
=
2
3
sin
2
sin 1
2
cos
2
2 1
u u u
o o
r
a
yy
(2.49b)

60
( ) stress Plane
zz
, 0 = o ( ) ( ) strain Plane
yy xx zz
, o o u o + = (2.49c)

(

|
.
|

\
|
=
2
3
cos
2
cos
2
sin
2
2 1
u u u
o t
r
a
xy
(2.49d)
0 (2.49e) = =
yz xz
t t
where is Poissons ratio, and and r are defined in polar coordinates, as shown in
Figure 2-29.



y
x
Crack tip
z
Figure 2-29: Definition of coordinate system ahead of a crack tip [20].


Examination of equation 2.49 reveals two significant results. The stress near the tip of a
crack is only dependent on two variables (r and ) and secondly, that it scales with the
remote stress and square root of the crack length. This scaling factor was termed the
stress intensity factor (K) given by:
a K t o = (2.50)

61
which, when substituted into equation 2.49, produces the following relationships:

(

|
.
|

\
|
=
2
3
sin
2
sin 1
2
cos
2
u u u
t
o
r
K
xx
(2.51a)

(

|
.
|

\
|
+ =
2
3
sin
2
sin 1
2
cos
2
u u u
t
o
r
K
yy
(2.51b)

(

=
2
3
cos
2
cos
2
sin
2
u u u
t
t
r
K
xy
(2.51c)
The stress intensity factor carries all the necessary information to describe the
stress field near the crack tip. It also provides a means of relating the stress field at the
crack tip to the applied far field stress used in design. The crack geometry and geometry
of the body containing the crack must also be considered. These geometric effects are
accounted for by inserting a coefficient Y:
a Y K t o = (2.52)
The geometric coefficient Y is the solution to a function that is dependent on the loading
configuration and body-crack geometry. These functions have been tabulated for many
configurations [65, 66]. Equation 2.52 is possibly the most recognized equation in
fracture mechanics and provides a tool to quantitatively describe a materials resistance to
fracture. The limitations of applying the stress intensity approach are addressed in the
following section.
The Griffith energy release rate (with units of J/m
2
or N/m) and the stress
intensity factor (with units of MN/m
3/2
or MPa-m
1/2
) have been introduced thus far. As
their names suggest, G
c
quantifies toughness in terms of energy and K
c
in terms of
magnitude of the stress field near the crack tip. Employing all the assumptions made for

62
each parameter separately, LEFM allows for the conversion of one toughness
measurement to the other in the following manner:
E E Stress Plane
E
E Strain Plane E G K
c c
=

= = ' ,
1
' , '
2
u
(2.53)
This relationship comes from the combination and simplification of equations 2.41 and
2.52.


2.4.3 Plane Stress vs. Plane Strain

In the field of fracture mechanics there is a strong emphasis given to whether a
component fails under plane-strain or plane-stress conditions. These conditions are
described by:
(2.54a)
(
(
(

=
0 0 0
0
0
yy yx
xy xx
Strain Plane c c
c c
c
(2.54b)
(
(
(

=
0 0 0
0
0
yy yx
xy xx
Stress Plane o o
o o
o
relative to the coordinate system shown in Figure 2-30 for a compact tension specimen
(C(T)).


63

Plane-Stress Plane-Strain
Figure 2-30: Schematic representation of thin (C)T sample in plane-stress and thick (C)T
sample in plane-strain.



The attention given to the topic of plane-strain vs. plane-stress conditions is well
deserved. It is observed that the apparent fracture toughness of a material varies with
specimen thickness, as shown in Figure 2-31, and that this variation is intimately related
to plane-stress and plane-strain conditions.


Figure 2-31: Experimentally observed dependence of stress intensity factor on specimen
thickness [69].



z
z
x
y
y
x
=0
=0 ZZ
ZZ
=0
=0 YZ
YZ
=0
=0 XZ
XZ

64
As the thickness of the material is increased, the apparent toughness decreases
until a point were it plateaus out at a constant value. At this point, the specimen is said to
be experiencing plane-strain conditions during fracture. This constant stress intensity
value is considered to be a material property: the fracture toughness of the material (K
C
).
Under these conditions, the critical stress intensity factor value is a function of only the
microstructure of the material. As the sample becomes thinner, geometric effects come
into play. At greater thicknesses the tip of the crack is under greater constraint. This is
because the plastic zone at the tip of the crack is embedded within a larger elastic region
that inhibits strain in the z-direction, producing a triaxial stress state at the tip of the crack
which results in higher stress levels. This is reflected in the planar appearance of the
fractured surface. As the sample becomes thinner it is said to be experiencing plane-stress
(or mixed) conditions and the constraint at the crack tip is reduced, lowering the stresses.
The lower stress level is a result of the plastic zone around the tip of the crack
approaching a free surface of the body which does not as effectively inhibit strain in the
z-direction and the fracture condition transitions from plane-strain to plane-stress. As this
occurs, the stress state at the tip of the crack evolves from triaxial to biaxial and the
fractured surface goes from being predominantly planar to forming shear lips. These two
features are, in actuality, the most relevant in determining if the measured stress intensity
factor is a valid K value. The general features of plane-strain and plane-stress fracture
are summarized in Table 2-8.



IC

65
Table 2-8: Summary of the general features of plane-strain and plane-stress fracture.
Plane-strain Plane-stress
Level of constraint at crack tip High Low
State of stress at crack front Triaxial Biaxial
Thickness of component (B) Thick Thin
Fractured surface morphology: (shear/planar) Low High
Fracture mechanics LEFM EPFM



The stress intensity factor is only a material property (the fracture toughness)
when the crack tip is under the highest possible constraint. This condition occurs when
the material is thick enough to generate a triaxial stress state around the tip of the crack
that is large enough to produce an adequate ratio of planar fractured surface to sheared
fractured surface. Beyond this point, the stress intensity factor becomes independent of
geometry and does not change.

2.4.4 Nonlinear Fracture Mechanics (J-Integral)

It is well known that not all components fail under elastic conditions and in a
brittle manner. This is particularly true for alloys that exhibit high ductility and relatively
low yield strengths. Attempting to collect valid fracture toughness data for design
purposes or to rank alloys relative to each other can be problematic because the
assumptions employed in LEFM are no longer valid. As depicted in Figure 2-31, if the
thickness of the material being fractured is increased to a size where the constraint at the
crack tip is maximized, the difficulty of characterizing the fracture toughness of a highly
ductile material is overcome. In practice, this is not as easily achieved as one may think.
Obtaining toughness values to comply with the criteria set forth by LEFM can, in certain
materials, require a specimen with length dimensions on the order of meters, applied load

66
capacities on the order of tons, and specialty gauges. A typical testing facility does not
have these capabilities and representative material of the necessary size is sometimes
unavailable.
Elastic-plastic fracture mechanics (EPFM, also referred to as non-linear fracture
mechanics) was developed out of LEFM in order to address the issue of the creation of a
plastic zone around the crack tip. One of the earliest and most substantial contributions to
EPFM was made by Rice [70-72] in 1968 when he used the mathematical frame work set
forth by Eshelby [73] to show that the energy in a region around the crack tip can be
described using a line integral. This relationship, termed the J-integral, is given by:

j ij ij ij
C
i
i
n T d w ds
x
u
T wdy J
ij
o c o
c
= = |
.
|

\
|
c
c
=
} }
, ,
0
(2.55)
where w is the strain energy density, T is a traction vector acting on the boundary, u is a
displacement vector, and ds is an increment of length along the contour C, as shown in
Figure 2-32.


Crack tip
Figure 2-32: Schematic representation of a line contour surrounding a crack tip [20].


67
Mathematically, the line integral of a closed contour must be equal to zero. What
this means physically, when applied to a cracked body, is that under an externally applied
load, the stresses that arise around the boundary are balanced. It is easily seen that J
Q
and
J
R
are boundaries at free surfaces and both are equal to zero. This means that since the
crack is stationary (dy=0) the sum of J
C
and J
C
must be equal in value and opposite in
sign. The contours J
C
and J
C
being of equal value reveals that the line integral is path
independent. In order for this analysis to hold true it is assumed that stress and strain are
both path-independent state functions (deformation theory of plasticity). This treats a
nonlinear elastic-plastic material as if it were non-linear elastic so that stress and strain
are uniquely defined at only one position.
The J-integral allows for the comparison of the change in potential energy (strain
energy being stored in the material) for an increment of crack extension (da along dy). It
has the same units as G (J/m
2
or N/m) and can be thought of as a nonlinear energy release
rate. When conditions exist that allow for LEFM to be used then:
'
2
E
K
G J
C
C C
= = (2.56)
As the plastic zone at the tip of the crack increases in size and maximum constraint is lost
this relationship becomes invalid.
Although the line integral form of J was first used, and its derivation was the
rigorous validation of the theory, it is rarely used in practice. The main reason for this is
the difficulty in monitoring a contour in a 3-D body being loaded. In practice J values are
calculated by using the load-displacement curve of a specimen loaded in displacement
control to determine the energy absorbed per increment of physical crack extension. A
graphical construct is then used to determine the critical value of J.

68
2.5 Relationship between Strength and Toughness

The inverse relationship between yield strength and fracture toughness, shown in
Figure 2-33, has long been recognized. The available combinations of these properties
for a given alloy are what designers frequently use to size components (modulus also
plays a role in many applications). The available strength-toughness combinations are
often a limiting factor adding an additional constraint to the design methodology. In this
section, the nature of this inverse relationship will be discussed from a mechanics point
of view. The processing of an alloy and resulting microstructure have a direct and
substantial influence on the toughness (as they do on the yield strength). The effect of
processing and microstructure on toughness will be discussed in the subsequent section.



Figure 2-33: Fracture toughness plotted against yield strength showing an inverse
relationship [74]. The limit lines represent the loci of combinations of maximum values
of K
IC
and yield strengths.

Yield Strength (MPa)
+ |
IC y
K , o
F
r
a
c
t
u
r
e

T
o
u
g
h
n
e
s
s

(
M
P
a

m
0
.
5
)


69
The two predominant mechanical/geometric factors that influence the inverse
strength-toughness relationship are the sharpness of the crack tip and size of the plastic
zone. Note that these factors are not independent of microstructure. It was previously
shown that a crack acts as a local stress riser, the magnitude of which is directly
dependent on the sharpness of the tip (equations 2.46, 2.47, and 2.48). It is known that the
minimum crack tip radius is determined by the atomic spacing of the crystal structure. In
reality, the crack tip radius is substantially larger. All fracture in metals is accompanied
by some degree of plastic deformation; regardless of the magnitude of the constraint at
the crack tip. The plastic flow at the crack front initially results in the blunting of the
crack before crack extension occurs. The stress concentration at the tip of a remotely
loaded crack is reduced by the blunting and the alloy's ability to plastically deform. The
magnitude of the crack tip blunting before crack extension occurs is dependent on the
yield strength of the material. The lower the yield strength and more ductile the alloy, the
larger the degree of blunting which results in a higher fracture toughness. The inverse is,
of course, also true.
The analysis of the stress fields ahead of the crack tip (equation 2.49) predicts that
at the tip of the crack the magnitude of the stress will be the largest. These relationships
were derived assuming a linear elastic isotropic material. A result of these assumptions is
that the magnitude of the stress at the crack tip is over estimated, as shown by the shaded
region in Figure 2-34. As the component is loaded, the stress at the tip of the crack
increases until a point is reached where the yield strength is exceeded. When this occurs
the stresses are relaxed and a plastic volume is formed. A result of the formation of the
plastic zone is that a region extending a distance r
p
is carrying less load then predicted.

70
This has a similar effect as increasing the crack length. Irwin [75] corrected for this
decreased load carrying capacity by defining an effective crack length:
y eff
r a a + = (2.74)
which is the actual crack length plus a distance r
y
which is equal to the distance in front
of the crack tip where the predicted stress, assuming perfectly elastic conditions, reaches
the yield stress. When the effective crack length is substituted in to equation 2.52 the
effective stress intensity value becomes:
( )
eff eff eff
a a Y K t o = (2.75)
which is larger in magnitude than the predicted value producing a greater toughness. It is
easily realized that the effective crack length scales with yield strength and that as yield
strength increases both the plastic zone size and toughness will decrease. Blunting and
plastic zone formation are interrelated, but the blunting effect manifests itself in the stress
concentration while the plastic zone formation affects the perceived crack length.



Figure 2-34: Predicted elastic stress near a crack tip and estimation of the actual elastic-
plastic stress distribution [58].

71
At first it may not be apparent why r
p
(elastic-plastic conditions) and r
y
(elastic
conditions) have different values. The plastic region formed must be large enough to
carry the load that would have been carried in a perfectly elastic body. In order to achieve
this, the plastic zone must extend a distance farther than that predicted by the elastic
stress analysis. It takes some amount of energy/work per unit volume to deform this
region. The size of this volume for a given loading condition and configuration is directly
related to the yield strength. As the yield strength decreases, the plastic zone size
increases, resulting in greater energy absorption and an increased toughness.

2.6 Metallurgical Aspects of Fracture

The fracture of metals is usually characterized as either being brittle or ductile in
nature. These two categories can be further subdivided into transgranular and
intergranular fracture, as shown in Figure 2-35.



72

Figure 2-35: Separation of fracture types and schematic representation of each [58, 76].
D
u
c
t
i
l
e

B
r
i
t
t
l
e

Intergranular Transgranular
GBDF
Intergranular Brittle
(Environmentally assisted)
Cleavage
Dimpled Rupture


In an idealized case of a high purity metal without any internal flaws, subjected to
uniaxial tension, fracture will occur by dislocation slip. That is, strain localization occurs
and the material necks. Then the reduced region continues to plastically deform until it is
drawn down to a point at which the material separates, as shown in Figure 2-36. This
type of fracture is never observed in engineering alloys.


73

Microscopic yielding
Ductile fracture Brittle fracture


Microscopic & macroscopic yielding
X
X
X
Pure metal
Figure 2-36: Tensile engineering stress-strain curve depicting brittle and ductile behavior.


As discussed in previous sections, obstacles are present in engineering alloys that
inhibit plastic flow producing higher strengths. These obstacles, along with alloy
contamination, act as local stress risers which create microscopic cracks and/or voids. In
materials that exhibit ductile failure, these stress risers are typically second phase
particles (oxide, constituents, precipitates, and dispersoids) that either delaminate or
fracture depending on the strain incompatibility and interface cohesion between the alloy
matrix and particle. The nucleation of a void is subsequently followed by growth and
finally void coalescence resulting in failure, as shown in Figure 2-37.



74



Void nucleation:
Resulting from either the
cracking of decohesion of
second phase particles.
Void growth:
Resulting from continual
increase of stress
Void coalescence:
Resulting from strain
localization between
voids along a plane.
Figure 2-37: Schematic process of ductile failure resulting in a fractured surface that
exhibits extensive plastic deformation and dimpled rupture [58].



When failure occurs by void nucleation, the growth and coalescence of the voids
are usually the controlling factors. The reason for this (regardless if it occurs in the
necked region of a tensile specimen or at the tip of a crack in a fracture specimen) is that
the triaxial stress state provides ample stress to nucleate a void. Once voids are present,
the material between them acts much like a tensile specimen, the size of which is
dependent on the void spacing. The closer the voids, the easier it is to coalescence and the

75
lower the toughness, as shown in Figure 2-38 relative to the orientations shown in
Figure 2-39.


Figure 2-38: Effect of constituent particles on 2XXX and 7XXX series Al alloys [77].



ND
RD
TD
Figure 2-39: Samples orientation relative to plate orientation [78].

76
It is apparent that as the alloys become cleaner and the volume fraction of
particles is reduced, the toughness increases. The effect of mechanical fibering is also
apparent. As a material is mechanically worked, the grains flow to accommodate the
macroscopic shape change. This also has the effect of stringing out any pre-existing
particles. In a cold rolling operation these particles are preferentially distributed in the
RD direction on the TD-RD plane. This non-uniformity in the microstructure is seen in
the anisotropic fracture toughness values. The specimen orientations in which the crack
propagates in either the RD direction or is parallel to the ND plane have the lower
toughness values. Mechanical processing can also damage particles and aid in the
nucleation of voids.
The fractured surface of a material that has failed by this void coalescence shows
excessive plastic deformation and a characteristic dimpled topology. The spacing of the
dimples has been shown to correlate with the particle spacing, as shown inn Figure 2-40
[79]. This type of ductile failure is referred to as dimpled rupture and usually occurs
transgranularly.


77

Figure 2-40: Inclusion spacing plotted against dimple size for age hardening Al alloys
[79].



A second type of ductile failure is grain boundary ductile fracture (GBDF) [76].
This is very similar to dimpled rupture except it is intergranular and has the visual
appearance of a facetted brittle failure at low magnification. At high magnification it has
a dimpled ruptured appearance because the fracture mechanism is essentially the same.
GBDF occurs when there is a large area fraction of second phase particles on grain
boundaries. This is common in precipitation hardened alloys due to the lower barrier to
nucleation at the grain boundaries. The process occurs in the same manner as dimpled
rupture except the grain boundaries create a path of closely spaced particles. GBDF
requires less work then transgranular ductile failure.
Brittle failure can also occur either intergranularly or transgranularly. This type of
fracture shows very little plasticity and has a facetted surface, as shown in Figure 2-41.
The intergranular method by which metals fracture in a brittle manner is cleavage. This

78
type of fracture is characterized by the rapid separation of crystallographic planes. In
order for cleavage to occur, plastic flow must be restricted. This type of fracture is not
seen in fcc alloys due to the high number of available slip systems. When cleavage does
take place, it occurs on the planes with the lowest packing density because it requires the
least number of bonds to be broke and the spacing between the planes is the greatest. The
crack is usually initiated by a local stress raiser in the form of a particle.



Figure 2-41: Micrograph of a cleavage fracture showing a facetted surface and no
ductility.



Figure 2-42: Micrograph of a intergranular brittle fracture showing a facetted surface and
grain structure.

79
Intergranular brittle fracture occurs when the grain boundaries become embrittled.
This is most often due to segregation at the grain boundaries or the effect of the service
environment on the alloy. The fractured surface shows no plasticity and an apparent grain
structure, as shown in Figure 2-42.


80
CHAPTER 3. MATERIAL

The purpose of this section is to give a very basic introduction to Al alloys and
tempers. The main alloy of interest in this study, AA2524 a higher purity version of
AA2024, will also be introduced.


3.1 Aluminum Alloys

Aluminum alloys are generally separated into two categories; heat treatable and
non-heat treatable, as shown in Table 3-1. The heat treatable alloys are those which are
capable of being precipitation hardened. Note that some Al-Li 8xxx series alloys are heat
treatable while other 8xxx series alloys are not. The high-strength aluminum alloys
(2XXX and 7XXX) all achieve their mechanical properties through some sequence of
precipitation reactions sometimes combined with thermal-mechanical processing. The
basic temper descriptions are shown in Table 3-2. These designations are often followed
by up to four numeric digits which provide more specific information about the
processing history of the material. Depending on, not only the type of processing, but
also the processing sequence, a wide range of mechanical properties can be produced for
a given alloy.






81
Table 3-1: Alloy series with majority allowing elements.
Series Alloying Elements Heat Treatable
1XXX >99.00% Al No
2XXX Cu Yes
3XXX Mn No
4XXX Si No
5XXX Mg No
6XXX Mg and Si Yes
7XXX Zn Yes
8XXX Other Yes/No
9XXX Unused N/A




Table 3-2: Basic temper designation and descriptions.
Temper Designation Discripton
F As-fabricated
O Annealed
W Soultion heat treated
H Strain hardened
T Thermally treated (percipitation heat treated)




An important factor to be considered when designing with heat treatable
aluminum alloys is quench sensitivity [80, 81]. Quench sensitivity, as the name implies,
is the sensitivity of an alloy's properties/performance to quenching parameters. An alloy's
quench sensitivity is of particular importance in structural application because it most
often reveals itself as a degradation in properties with increased product thickness. The
quench sensitivity (improved over that of 7050) of Alcoa's most modern 7
th
generation
structural alloy, 7085-T7452 [82], is illustrated in Figure 3-1, Figure 3-2, and Figure
3-3. As the thickness of the hand forgings increases the effective quench rate decreases as
a result of there being more heat to be pulled out a think cross section compared to a
thinner one. The thicker the product form, the longer it takes the center of the material to
reach the temperature of the quench tank. The effects of this slower quench rate are that:
(1) coarse non-strengthening precipitates form during the quench, (2) there is less solute

82
available to precipitate the strengthening phases during subsequent aging do to it being
lost to the non-strengthening precipitates, and (3) there is a persistent through-thickness
macrostructure gradient. The negative impact on strength is a consequence of slower
aging kinetics and less available solute. The fall in toughness with increased thickness is
a result of the coarse non-strengthening precipitates that form (usually on grain
boundaries) during the quench. Quench sensitivity also has a significant impact on
corrosion properties due to galvanic coupling within the microstructure. Aluminum alloys
of certain product forms and tempers are prohibited on Naval Aviation Systems above
specified thicknesses because of their susceptibility to corrosion and the increased
maintenance required.
It should also be noted that cladding is a common practice in the aluminum
industry. Cladding is when commercial purity Al is roll-bonded to an alloy in order to
provide a corrosion barrier.


83

Figure 3-1: Tensile yield strength plotted against product thickness for 7085-T7452 and
7050-T7452 S-Basis. MMPDS MIN for 7085 is a tentative A-Basis [82].



Figure 3-2: Ultimate tensile strength plotted against product thickness for 7085-T7452
and 7050-T7452 S-Basis. MMPDS MIN for 7085 is a tentative A-Basis [82].

84
Figure 3-3: Fracture toughness in the L-T orientation plotted against product thickness for
7085-T7452. S-Basis minimums are tentative [82].


Quench sensitivity within an alloy system is most strongly dependent on the
solute level and ratio of alloying elements. The quench sensitivity between alloy systems
is more dependent on the dispersoids present. The dispersoids in 7xxx series alloys form
coherent interfaces which do not easily facilitate heterogeneous nucleation during the
quench. The 2xxx series alloys form incoherent interfaces which are effective
heterogeneous nucleation sites. The incoherent interfaces also generate dislocations
during the quench which further facilitates precipitation during quenching. Quench
sensitivity will be revisited in Chapter 4.



85
3.1.1 AA2524

Aluminum alloy 2524 is a heat treatable 2XXX series alloy with Cu and Mg as its
major alloying elements, as shown in Table 3-3. Alloy 2524 is a modification to the
original alloy 2024. The second digit in the alloy number indicates that it is the fifth
modification. Modifications to the original alloy most often reflect a lower impurity
(usually Si and Fe) requirement.

Table 3-3: Chemical composition by weight% for original alloy 2024 and higher purity
version 2524.
Aluminum (Al) Chromium (Cr) Copper (Cu) Iron (Fe) Magnesium (Mg) Manganese (Mn)
90.7 - 94.7 % 0.100 % 3.80 - 4.90 % 0.500 % 1.20 - 1.80 % 0.300 - 0.900 %
---- Other, each Other, total Silicon (Si) Titanium (Ti) Zinc (Zn)
0.0500 % 0.150 % 0.500 % 0.150 % 0.250 %
Aluminum (Al) Chromium (Cr) Copper (Cu) Iron (Fe) Magnesium (Mg) Manganese (Mn)
92.5 - 94.4 % 0.0500 % 4.00 - 4.50 % 0.120 % 1.20 - 1.60 % 0.450 - 0.700 %
---- Other, each Other, total Silicon (Si) Titanium (Ti) Zinc (Zn)
--- 0.0500 % 0.150 % 0.0600 % 0.100 % 0.150 %
AA2024
AA2524


The major compositional differences between AA2524 and the original version are:
a) Reduced chromium. This has the effect of slightly reducing the quench sensitivity
and reducing the temperature at which recrystallization will occur. Altering the
energy barrier for recrystallization has no effect on the current project, while
lessening the quench sensitivity is beneficial. This allows for a more effective
quench resulting in higher levels of solute remaining in solid solution.
b) Iron levels are greatly reduced. Fe is essentially an insoluble contaminant that is
too costly to refine completely out of the melt. It provides no metallurgical benefit

86
to any Al alloy system. There are some limited processing applications where Fe
provides a beneficial "die cleaning/release" function.
c) Manganese level has been increased. Manganese has many functions. It readily
combines with Fe and Si preventing them from forming less favorable phases
and/or tying up the main alloying elements (Cu and Mg). It is also an effective
element for solution hardening. The drawback is that it increases the quench
sensitivity.
d) In 2XXX series alloys, Si is much like Fe and has been substantially reduced.
Unlike Fe, the Si can affect the aging process although it is usually in a
detrimental way.
e) Titanium and Zinc have both been reduced. These elements are added as grain
refiners during casting. They also form dispersoids and can act as nucleation sites
for relatively large precipitates that do not strengthen as effectively as smaller
ones.

The 2524 alloy is often used in the aerospace industry, as well as many others, in
either the naturally aged or artificially aged condition. The precipitation sequence is as
follows:
GPZ S S
with being a super saturated solid solution present after quenching, followed by fully
coherent GP zones, which upon exposure to elevated temperature progress into the
metastable semi-coherent S phase, which in turn transfers into the non-coherent stable S
phase (Al
2
CuMg). The GP zones form rods that are approximately two atomic layers

87
thick and 40nm in diameter. Both the S and S phases form laths which vary in size. In
the peak aged condition all phases are typically present; although, the GP zones have
been partially dissolved and transformed into a mix of both S and S. Precipitation
hardening, as well as other hardening methods, were discussed earlier in greater detail.


88
CHAPTER 4. EFFECT OF GP ZONE FORMATION ON STRAIN HARDENING
RATE


High strength Al alloys (2xxx and 7xxx) achieve their properties through
precipitation hardening. These alloys have been specifically designed to respond to this
strengthening method and the process is routinely carried out successfully. Precipitation
hardening is a desirable strengthening method, not only because it is effective, but
because it allows for components to be shaped at relatively low strengths (O-temper)
before being heat treated to their service strength. Precipitation hardening can also be
easily performed on complicated shapes. An unavoidable consequence of artificial aging
is that while the strengthening precipitates are forming in the interior of the grains there
are also less coherent precipitates forming at the interfaces between adjacent grains (grain
boundaries). This is a consequence of the relationship that exists between
crystallographic coherency and the barrier to nucleation [57].
Grain boundary precipitates do not contribute to the static properties of an alloy,
but are known to degrade the fracture toughness of the material [76, 83-86]. This
undesirable microstructure feature can be avoided if other strengthening methods are
employed. Strain hardening has been largely ignored as a viable strengthening option in
Al based alloys due to their high stacking fault energy (SFE) which promotes dynamic
recovery. The exception being the beverage-can industry, which does strain harden high
strength Al alloys [87]. Although, this is not a high reliability or structural application.
The gauge thickness of beverage-cans is also relatively thin so the issues of quench
sensitivity and residual stress are not encountered.

89
It has been shown by Doherty et al. [5, 6, 92] that potent strain hardening rates
(SHR, d/d) can be achieved in heat treatable Al alloys provided that the solute is not
precipitated out of the matrix. It is clearly exhibited in Figure 4-1 and Figure 4-2 that as
the amount of solute is increased, not only is the flow stress at equivalent strains
increased but the instantaneous slope of the curve is increased as well. As the artificial
aging process progresses, and precipitates form, the matrix becomes depleted of solute
and the SHR is diminished.

y = 90x
0.3
y = 269x
0.19
y = 372x
0.17
0
100
200
300
400
500
600
0 0.5 1 1.5 2 2.5 3
Rolling Strain
T
e
n
s
i
l
e

Y
i
e
l
d

S
t
r
e
s
s

(
M
P
a
)
0
10
20
30
40
50
60
70
80
T
e
n
s
i
l
e

Y
i
e
l
d

S
t
r
e
s
s

(
K
s
i
)
3Cu-1Mg 2Cu-0.5Mg
y = 448x
0.19
1.5Cu-0.5Mg 1Cu-0.5Mg
Al 99.99%
y = 317x
0.19

Figure 4-1: Tensile yield strength plotted against rolling strain for Al alloys with
increasing amounts of solute [5]. Each data set has been fitted to an exponential equation.



90
y = 27.783x
-0.7296
y = 52.947x
-0.8467
y = 62.395x
-0.8467
y = 65.598x
-0.8681
0
20
40
60
80
100
120
140
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Rolling Strain
S
H
R

(
M
P
a
)
0
5
10
15
20
S
H
R

(
K
s
i
)
3Cu 0.5Mg y = 65.435x
-0.8895
2Cu 0.5Mg
1.5Cu 0.5Mg
1Cu 0.5Mg
Al 99.99%

Figure 4-2: Stain Hardening Rate plotted against rolling strain for Al alloys with
increasing amounts of solute [5]. Each data set has been fitted to an exponential equation.



Unfortunately, when the work carried out by Doherty et al. [5, 6] was performed
there was little attention paid to the formation of GP zones and their effect on strain
hardening. It is assumed that the material was processed in the fully naturally aged
condition (T4-temper) due to the practical difficulties of doing otherwise but this is
uncertain. It is not unreasonable to expect that a gliding dislocation will interact with the
stress field associated with a solute atom differently than that associated with a semi-
coherent solute cluster. The interaction of these stress fields impedes dislocation flow;
increasing, not only, the yield strength of the alloy but also the SHR.

91
The above discussion has prompted the following questions to be posed: if solute
enhanced strain hardening (SESH) of high strength Al is a possible processing route for
achieving improved combinations of strength and toughness:
1) What is the optimum condition to deform the material in;
2) Is the strain hardening of heat treatable Al alloys more effective with solute in
solution or clustered into GP zones;
3) What is the effect of a high dislocation density on the subsequent formation of GP
zones?

4.1 Natural Aging

The aging process in heat treatable aluminum alloys is subdivided into natural
aging and artificial aging. The natural aging process is the room temperature
transformation of an aluminum matrix (also occurs in Ag and Zn based alloys) super
saturated with solute into a matrix containing fully coherent solute clusters known as
Guinier-Preston (GP) zones. The duration of the natural aging process is on the order of
days. Natural aging is initiated upon rapid quenching after solution heat treating and
terminates when the equilibrium concentration of vacancies is reached. The changes in
physical properties associated with natural aging also stabilize when the process
terminates.
Artificial aging is a classical nucleation and growth transformation. There is a
barrier to nucleation that must be overcome which requires elevated temperatures for the
transformation to occur. The elevated temperature is also necessary to facilitate the long
range diffusion of solute to precipitates. As time and temperature are increased, the

92
precipitates will simultaneously evolve into more stable structures and undergo Ostwald
ripening. The changes in physical properties associated with artificial aging are more
pronounced then natural aging and will continue to evolve at temperature.
The current Chapter of this document concentrates on the natural aging process
which will be described in more detail in subsequent sections. Before proceeding with the
discussion it is important to distinguish between natural aging and artificial aging which
are fundamentally different microstructure transformations. Artificial aging is a
discontinuous reaction where localized nuclei of a new phase form (Gibbs type I
reaction) where natural aging is a continuous reaction that is delocalized in the structure
(Gibbs type II reaction).

4.1.1 Solution Heat Treating

A necessary condition for the effective strengthening of a heat treatable Al alloy,
regardless of the strengthening method (SESH, natural aging, or artificial aging), is
solutionizing. During the solution heat treating (SHT) process, the alloy is taken up to an
elevated temperature in order to dissolve pre-existing precipitates and put the solute
atoms back into solution. The material is then rapidly brought back to room temperature
locking the solute in solution. If the solution heat treatment is performed correctly, the
result is a super saturated solid solution at room temperature.
The two most critical parameters in the SHT process are: (1) the solutionizing
temperature and (2) the rate of cooling. From a metallurgical point of view, it is most
beneficial to SHT at the highest possible temperature followed by the fastest cooling rate
achievable. As the SHT temperature increases, so will the solubility level, allowing for

93
more of the solute to be put into solution. The upper limit of the SHT temperature range
is bounded by the solidus temperature which is usually the eutectic melting point. Above
this upper temperature limit the material will partially change state resulting in poor
mechanical properties regardless of subsequent processing. After the solute has gone into
solution the material must be brought back down to room temperature. The rate at which
this temperature change occurs is critically important being that solubility is a function of
temperature. As the temperature is decreased, the solute will precipitate out of solution
forming various phases. At these elevated temperatures heterogeneous nucleation at grain
boundaries and inclusions is most likely to occur, not only making the SHT less effective
but negatively impacting the mechanical properties of the material. In order to retain the
solute in solution and prevent the precipitation of metastable phases the material must be
quenched as rapidly as possible. All phase transformations in Al based alloys are
diffusion controlled and while solubility is a function of temperature, so is the rate of
diffusion. If the material is cooled from the SHT temperature to room temperature at a
rate faster than the atoms can rearrange themselves into a lower energy structure the high
temperature microstructure is preserved. This results in an aluminum matrix super
saturated with solute atoms.

4.1.2 Excess Vacancies

In the previous section, the effects of the solutionizing temperature and cooling
rate on the amount of solute retained in solution at room temperature were discussed.
These two processing variables also impact another important microstructure feature,
lattice site vacancies. The equilibrium vacancy concentration is given by:

94

RT H
o
v f
e
n
n

= (4.1)
where n
v
is the number of vacancies, n
o
is the number of atoms, H
f
is the heat of
formation, R is the gas constant, and T is the temperature. As the SHT temperature
increases, the vacancy concentration increases exponentially. When the material is
quenched not only is the solute retained in solution, but the vacancy concentration at the
SHT temperature is as well. This results in a non-equilibrium (excess) concentration of
vacancies at room temperature. Relative to atoms, vacancies are highly mobile. The jump
rate of a vacancy is given by:

RT H
v
m
Ae r

= (4. 2)
where A is an empirically determined constant and H
m
is activation enthalpy. Vacancy
motion occurs by atoms jumping into their lattice site displacing the vacancy. This is the
same mechanisms by which atoms migrate. With this in mind, the probability of a
vacancy having an atom in a neighboring lattice site is significantly higher then the
opposite situation. The jump rate of a typical atom is given by:

RT H
o
v
a
m
Ae
n
n
r

= (4.3)
where the quantity of n
v
/n
o
is always much less then one. The mobility of the vacancies
allows the microstructure to achieve its equilibrium vacancy concentration over time (on
the order of days in Al based alloys) at room temperature. This occurs by the vacancies
migrating to sinks [93] where they are consumed. Microstructure features that consume
vacancies are dislocations (static, gliding, and walls) and interfaces (HAGB, LAGB,
phase boundaries, and free surfaces).

95
Vacancies can also be created by the bombardment of high energy particles and
the interaction of gliding dislocations. The first of these vacancy creation mechanisms is
of little importance for Al based alloys because their melting point excludes them from
most applications where they would be exposed to such an environment. The second
mechanism frequently occurs during the industrial processing of Aluminum. The
intersection of gliding screw dislocations results in the formation of non-conservative
jogs. The newly created jogs have the form of edge dislocations. In order for the original
screw dislocation to continue to flow in must drag the edge portion with it. The dragging
of the edge dislocation produces vacancies.

4.1.3 Guinier-Preston Zone Formation

The process of SHT produces a metastable microstructure that is super saturated
with solute atoms and contains a non-equilibrium concentration of quenched-in
vacancies. The microstructure subsequently lowers its energy level by a process termed
natural aging. The natural aging process takes place at room temperature and goes to
completion within the time frame of days in the 2xxx series. This time frame depends on
alloy composition and SHT parameters. The fully naturally aged condition is referred to
as the T4-temper.
The natural aging process is a Gibbs type II transformation in which the solute
atoms cluster together producing fully coherent disks or rods known as GP zones [94,
95]. These zones are typically two atomic layers thick and 10nm in length/diameter with
an approximate spacing of 10nm. The formation of GP zones is energetically favorable at
room temperature because the driving force produced by the super saturation is greater

96
then the increased interfacial energy, which is minimized by the coherency of the
zone/matrix crystal structures, and increased strain energy, which is minimized by the
zones forming such that the large side of the aspect ratio is perpendicular to the
elastically soft <100> crystal directions. Although the formation of the GP zones is
thermodynamically favorable, the bulk diffusion rate of the solute in Al is so small that
the time frame for them to form would be orders of magnitude longer then the days
observed.
In aluminum based alloys all possible phase transformations take place by
diffusion, the specific mechanism being vacancy diffusion. Equation 4.3 shows that the
jump rate of an atom is directly dependent on the equilibrium vacancy concentration,
equation 4.1. However, after SHT and during GP zone formation, the structure does not
contain the equilibrium concentration of vacancies but rather a much larger vacancy
concentration (estimated to be 145x larger in the current study). This greatly increased
vacancy concentration increases the mobility of the solute atoms, allowing them to
cluster and form GP zones [96]. As time elapses and the vacancy concentration decreases
to its equilibrium level, the rate of GP zone formation decreases until it is eventually
halted.
The natural aging process is difficult to observe due to its dynamic nature and the
physical size of the GP zones. The instrument most commonly used to directly
investigate these structures is the transmission electron microscope (TEM). It is often
much more practical to indirectly observe the natural aging process. Because the natural
aging process is a gradual change of the microstructure it impacts physical properties of
the material. The two most commonly tracked physical changes are: (1) eddy current

97
electrical conductivity, which decreases as GP zones form and more effectively scatter
electrons, and (2) macrohardness, which is proportional to yield strength and increases
due to the interference of GP zones with dislocation flow. All physical/mechanical
property changes exhaust as GP zone formation ends.
Natural aging should not be confused with artificial aging which is a series of
Gibbs type I transformations that occur at elevated temperature. These are two distinctly
different aging processes. Artificial aging is used to produce the T6, T7, T8, and T9
tempers.

4.1.4 Residual Stress

It was previously stated that the SHT process is most effective when performed at
the highest possible temperature and fastest cooling rate. While this is true
metallurgically, there are negative consequences to following this procedure. An
unintentional byproduct of this rapid change in work-piece temperature is that a sustained
stress (residual stress) is setup in the material [97-103]. The residual stress is a result of
non-uniform cooling. When the work-piece is taken up to the temperature range required
to put the solute back into solution there is a geometric change, the magnitude and nature
of which depends on the coefficient of thermal expansion (CTE), crystallographic
texture, and material constraint depending on the bulk shape. When the material is
quenched into room temperature water the heat is rapidly pulled out of the surface while
the center initially remains hot. This creates a steep temperature gradient through the
thickness of the material. The significance of this temperature gradient is that it results in
the material contracting at different rates through the thickness. The outer most layer

98
initially cools and contracts while the still hot inner material resists the thermal
contraction. Given that yield strength is a function of temperature, the still hot inner
material does not resist plastic deformation as effectively as it would at room
temperature. As the work piece continues to cool, the inner material reaches room
temperature and contracts, which is then more strongly (due to the outer shell already
being at room temperature) resisted by the outer shell which has already achieved its
dimensionally stable condition while the inside was at temperature. This non-uniform
cooling and contraction produces a compressive stress at the surface and tensile stress in
the interior, as shown in Figure 4-3. Although it may seam obvious, it is worth stating
that regardless of the magnitude and state of the residual stress it must be balanced. The
magnitude of the residual stress is dependent on the steepness of the temperature gradient
and the yield strength of the material. The state of the residual stress is dependent on 3-D
shape of the work-piece and the uniformity of the temperature gradient. It should be
noted that, although this discussion is concentrating on thermally induced residual stress,
residual stresses can also be induced through machining, deformation processes, surface
treatments, and while in-service.


99
Solution Heat
Treat Temp.
ND
Thermally expanded state
RD

Figure 4-3: Schematic representation of the non-uniform cooling that results when a
work-piece is quenched into room temperature water from SHT temperature and the
subsequent residual stress from the non-uniform contraction.


The magnitude and nature of the sustained stress is almost always unknown which
makes predicting its impact to performance difficult. An unfortunate consequence of the
inability to adequately predict the impact of residual stress on material performance is
that its presence is ignored by a large portion of the design community. The existence of
this present, but unaccounted, for stress often reveals itself though warping/distortion
during quenching and/or machining along with fit-up problems during assembly. When
components with residual stress are put into service the material will often exhibit
decreased fracture toughness, degraded fatigue performance, and unforeseen stress
corrosion issues.
There are two approaches to dealing with the presence of a residual stress;
management and relief. The management of residual stress addresses the producibility of
a component and not the stress its self. This is usually done by developing elaborate
Outside cools and contracts
while the still hot softer
inside resists
Rapid Quench
(water)
Inside cools and contracts
Outside resists contraction
Room Temp.
Tensile Stress Compressive

100
machining procedures to balance the stresses as material is removed, setting force limits
during part fit-up, and over sizing parts. The danger in choosing to manage residual stress
is that, although the dimensional stability issues have been overcome, the magnitude of
the stationed stress is still unknown and may not have been reduced at all. Stress relieving
addresses the magnitude of the sustained stress. Depending on the alloy composition and
product form there are various methods for stress relieving.
Typically, the stress relieving of metallic components falls into two basic
categories, thermal and mechanical. Both of these stress relieving methods are effective,
but limited, in application. Thermal stress relieving involves increasing the temperature
of the work-piece until the yield strength is exceeded by the residual stress. This manner
of stress relieving is not possible in heat treatable aluminum alloys because the
temperatures required for it to be effective result in the precipitation of metastable phases
and the undoing of the SHT. Mechanical stress relieving is an effective method for stress
relieving high strength aluminum, although it is only effective on certain product forms,
such as plate and rod, and is limited by product thicknesses. From a practical standpoint
mechanical stress relieving is most effective when the product form is uniform in
thickness and the state of the residual stress can be approximated as symmetric.
Mechanical stress relieving is not as effective on product forms such as high strength
aluminum forgings or castings because both the stress state and work-piece geometry are
much more complex.
It should be noted that there are other stress relieving methods that involve
interrupted quenching, precipitate reduction processes, high frequency vibration, and
cryogenic treatments. These are not widely used industrial processes and the

101
effectiveness of many of them is not agreed upon. In the author's opinion, the
discrepancies in the literature regarding the effectiveness of these more novel stress
relieving processes is a result of both not recognizing the importance of the source of the
residual stress and emphasizing only the magnitude of the stress while ignoring it's state.

4.2 Strain Path Change

The effect of a change in strain path on the uniform ductility of metals has long
been recognized in the stamping, deep drawing, and sheet forming industries. It was
shown by Considere [104] that strain localization is dependent on the rate of increasing
flow stress and decreasing SHR. When the flow stress becomes equivalent to the SHR
strain localization occurs, as described by:
n n
d
d
u
= = = c
c
o
c
o
o (4.3)
where
u
is true uniform strain and n is the strain hardening exponent from the Hollomon
equation (equ. 4.4). Abruptly altering the strain path has been shown to affect both the
yield point and SHR of the second deformation mode relative to the first [105-108].
Typically, the second strain path exhibits either a transient drop in SHR in conjunction
with a slightly increased yield strength or inversely a transient increase in SHR in
conjunction with a decrease in yield strength, relative to if the initial strain path was
continued. These observations between yield strength and SHR are not universally
observed. Transient changes in SHR have been reported in the absence of a change in
yield strength. There are also discrepancies in the literature regarding the effect of the
amount of initial strain on the magnitude of the change in SHR, although it is agreed that

102
there is a minimum amount of strain that needs to be imposed to produce the phenomena.
The effect of strain path change on SHR in fcc metals has most successfully been related
to SFE. As the SFE of an alloy decreases, the level of uniform elongation increases from
being below what is observed if the strain path is not changed to above.
A preliminary review of aluminum based alloys supports the trend that alloys with
high SFE show transient decreases in SHR and an earlier onset of strain localization. The
majority of the Al alloys that were used to investigate the effect of strain path change
were either high purity, commercial purity (CP), or annealed to the O-temper. There was
much less information available on higher solute alloys in the T4-temper or under aged
condition. This is not surprising since forming operations for heat-treatable alloys are
typically done in the fully softened O-temper and higher purity alloys are commonly used
for investigative purposes to decrease variables. Studies that have been conducted on Al
alloys with higher levels of un-precipitated solute [105, 106] exhibited substantially
different results. Wilson et al.[105] compared the response of both a CP aluminum and an
Al-Cu-Mg (2014) alloy subjected to a tensile strain path change as shown in Figure 4-4.



Figure 4-4: Effect of pre-strain followed by followed by a 90 change in direction of
uniaxial tension on the relationship of SHR and flow stress plotted against total effective
strain for (a) CP Aluminum and (b) an Al-Cu-Mg alloy [105].

103
After being pulled in tension, rotated 90, and pulled in uniaxial tension again the
CP material exhibits the typical response, producing an initially higher yield point and a
transient decrease in SHR that, as deformation continues, coincides with the single strain
path data. The higher solute alloy behaves in the manner predicted for low SFE alloys
(which it is not). The Al-Cu-Mg alloy has a lower initial yield strength when the loading
direction is changed and exhibits a transient increase in SHR.
Observations of the dislocation structures produced in the various strain path
change experiments conducted on Al based alloy provides greater insight into explaining
the discrepancies observed in alloys with higher levels of un-precipitated solute. In the
experiments where TEM analyses was performed [107] it was reported that well defined
dislocation cell structures were formed by the initial deformation. This is in agreement
with what was reported by Doherty et al. [6] for low solute alloys and is to be expected
for high SFE material in general. The well recovered dislocation cell structure is a result
of uninhibited dynamic recovery. When the deformation mode was changed, it was
observed that this preexisting sub-grain structure was first destroyed and a new sub-grain
structure was subsequently formed. The disruption of the preexisting dislocation cell
structure is reported to coincide with the decrease in SHR. These results fit the
experimental observation that there is a minimum amount of deformation that needs to be
imposed in the first deformation mode to see a transient effect in SHR in the second
mode because the initial cell structure has to be created before it can be subsequently
destroyed. The effect of SFE also fits the observation that the decreased SHR is a result
of the disruption of the cell structure because only high SFE materials form this type of
dislocation structure.

104
In the work done by [6, 107] the dislocation structures produced in low solute and
high solute alloys were described. Although the high solute alloys had a high SFE, the
deformation microstructures did not form dislocation cell walls but rather uniform
dislocation tangles throughout the microstructure. The lack of a cell structure to be
disrupted may explain why the higher solute alloys do not exhibit the reduced SHR;
although it does not explain why they do exhibit an increased SHR. It is reasonable to
expect that a change in strain path would necessitate activating dislocations on slip planes
that were previously inactive. These newly activated dislocations will experience the long
range stress fields of the previously activated dislocations and also intersect them,
effectively increasing the SHR temporarily.

4.3 Processing and Experimental Method

Alloy AA2524 was selected for this study because it would both effectively strain
harden and naturally age. The material was received form Alcoa in the form of an Alclad
plate in the F-temper. The Cladding was machined off and three coupons with
dimensions of approximately 152mm x 39mm x 10mm (RD x TD x ND) were removed
from the as-received plate for subsequent processing.

4.3.1 Heat Treating

Each of the three coupons were heat treated separately in order to better the
control processing conditions. The tight control of process variables allowed for the
relaxation of some of the more conservative guidelines specified in industrial standard

105
AMS-2770H [109]. The material was solution heat treated at a coupon temperature of
520-525C for 90min (the range in temperature is a result of multiple thermocouple
readings in contact with the coupon). The increased solutionizing temperature results in a
greater fraction of solute going into solution. At the completion of the solution heat
treatment, the material was water quenched, rather then quenched into a synthetic
solution (producing a slower and more uniform quench), to retain the maximum amount
of solute in solution at the cost of an increased residual stress. The time elapse between
the opening of the furnace door and immersion of the coupon into the room temperature
water was less than 5 sec. The quench tank contained ~30 gallons of water to prevent the
bath temperature from raising during the quench.
Once the coupon reached room temperature it was thoroughly dried and
immediately placed into liquid nitrogen (LN) to start the uphill quenching process
[110-116]. The time elapsed between the opening of the furnace door and the immersion
into the LN was approximately 60 sec. The material remained in the LN for 45min to
insure that the full thickness had stabilized at the subzero temperature (-196C). During
this time the kinetics of the natural aging process are slowed to a non-detectable point
[117]. The material was then uphill quenched into boiling water with a transfer time of
approximately 1 second and soaked for 5 seconds before once again being quenched into
room temperature water to complete the process, as shown in Figure 4-5 and Figure 4-6.
The end of the second room temperature water quench is taken to be the zero-time (the
start of the natural aging process) and the material is considered to be in the W-temper.
The uphill quenching process is believed to effectively reduce the residual stress
produced by water quenching. If this thermo-mechanical stress relief step is not

106
performed, subsequent deformation processing can not be carried out (section 4.4.1.1.
provides a description of the mid-plane cracking that occurred upon light rolling if only
water quenching was performed).



Figure 4-5: Flow chart depicting the processing steps involved in solution heat treating
(SHT) and the uphill quenching procedure.



-250
-150
-50
50
150
250
350
450
550
0 20 40 60 80 100 120 140 160
Time (min)
T
e
m
p
e
r
a
t
u
r
e

(
C
)
-418
-218
-18
182
382
582
782
982
T
e
m
p
e
r
a
t
u
r
e

(
F
)
SHT (250-525C f or 90min)
Dry Sample (~1min)
LN Soak (90min)
Boiling Water Quench (5sec)
Water Quench
Room Temp. (22C)

Figure 4-6: Temperature profile associated with the processing steps depicted in Figure
4-5.





Water quench
(room temp.)
Soaked in liquid
nitrogen
Boiling water
quench
Water quench
(room temp.)
Uphill Quench
Solution Heat
Treat

107
4.3.2 Deformation Processing - Cold Rolling

Each of the coupons were subjected to multiple cold rolling reductions utilizing
an electrically driven Stanat rolling mill with 150mm rollers. The rolling direction was
chosen such that it coincided with the previous rolling direction of the processes
performed at Alcoa (facility located in Bettendorf, IA). The first of the three coupons
processed was allowed to fully age to the stable T4 condition. The Rockwell B hardness
and eddy current electrical conductivity was tracked over this time period. Hardness and
eddy current electrical conductivity was performed in accordance with [118] and [119].
The coupon was then cold rolled in the T4 condition and specimens were removed from
the parent coupon at increasing rolling strains up to a true strain of 0.41. The hardness
and conductivity of each specimen was tracked post-deformation for an additional 22h
(almost twice the time taken to reach the T4 condition). The same procedures were
followed for the additional two coupons with the exception that the rolling was
preformed in the half-hardened condition (55min of natural aging) and in the W condition
(the zero-time immediately following the uphill quench). This procedure is schematically
represented by the flow chart in Figure 4-7.


108
Solution Heat
Treat
Water Quench
Uphill Quench
Cold Roll in the
W-Tem
Cold Roll at
intermediate time
Cold Roll in the
T4-Tem per per
= 0, 0.1, 0.2,
0.3, and 0.4
= 0, 0.15, 0.2,
and 0.4
= 0, 0.1, 0.2,
0.3, and 0.4
Compression
Testin
Compression
Testin
Compression
Testin g g g
Monitor HRB and EC
Data Analyses

Figure 4-7: Flow chart depicting the overall experimental procedure with the deformation
process branched into three paths representing different amounts of GP zone formation.


4.3.3 Mechanical Testing

Compression testing was performed on each specimen in order to determine the
yield strength and SHR at each natural aging condition and rolling reduction. The
compression testing method provided in [120] was used as a general guide. The testing
was carried out on cylindrical samples (diameter of 4mm and height of 6mm) taken from
each specimen such that the compression direction corresponded to the rolling direction.
The friction between the loading platen and sample interface was minimized by polishing
the bearing surfaces of the tooling to a mirror finish, employing the use of dry graphite

109
lubricant on the platen/sample interface, and by unloading the sample to 67N every
0.5mm of displacement to relax the forces constraining the radial expansion of the sample
at the interface. The compliance of the test system was eliminated by designing custom
tool-steel platens that allowed for clip gauges to be mounted between the bearing surfaces
at 120 intervals, as shown in Figure 4-8. The use of three clip gauges verified that the
compression faces remain parallel and vertically aligned throughout the test. The errors
resulting from the compression of the lubricant were diminished by preloading each
sample in its elastic region to even out any non-uniformity in the graphite coating, take-
up any initial non-recoverable displacement in the graphite, and ensure that the sample is
seated in the lubricant. The compressive nature of the lubricant/system was also
characterized by loading the system with a representative cylindrical tool-steel sample
and fitting a function to the load-displacement data so that the system compliance could
be subtracted from the experimental data before analysis.


Figure 4-8: Tooling using for compression testing.



110
4.4 Experimental Results

4.4.1 Heat Treating

The information contained in this section was not included in the Experimental
Procedure section of this document because it details processing routes that were deemed
unacceptable and were not used to produce test samples. However, it was the study of the
quenching rate that lead to the understanding of the mid-plane cracking during rolling
and the nontraditional processing route of uphill quenching. The experimental data
presented in this section will also be analyzed in the discussion section.

4.4.1.1 Quench Rate

Quenching into room temperature water is a very effective and common method
of rapidly cooling a work piece during heat treating. When the solutionizing was carried
out utilizing room temperature water as the quenchent it was not possible to subsequently
cold roll the material. Within the first three passes through the rolling mill, in-plane
cracking was observed near the mid-plane of the coupon, as shown in Figure 4-9. This
cracking is not a strain localization or shear banding event.



Figure 4-9: AA2524 plate that was solution heat treated and water quenched exhibiting
mid-plane cracking after light cold rolling.

111
The source of the cracking, regardless of the root cause (residual stress or quench
cracks), is directly linked to the rapid quench rate produced by the water. In order to
process the material without cracking it was necessary to reduce the quenching rate. This
was achieved by using a synthetic type I [121-124] polymer solution (Aqua-Quench)
supplied by Houghton International which provided both a slower and more uniform
cooling rate. The use of this synthetic quenchent facilitated the successful processing of
the material, allowing for cold rolling strains of 1.5 and higher to be imposed.
The natural aging profiles for alloy 2524 in both the water and polymer quenched
condition are shown in Figure 4-10, Figure 4-11, and Table 4-1. Examination of the
aging profiles quickly reveals that the material produced using the polymer quenchent is
not equivalent to that produced by water quenching.

























112
40
45
50
55
60
65
70
75
80
0 10 20 30 40 50 60 70 80 9
Time (hr)
H
a
r
d
n
e
s
s

(
H
R
B
)
0
Water Quench
Polymer Quench
T4 (Water Quench)
W (Water Quench)
T4 (Polymer Quench)
W (Polymer Quench)

Figure 4-10: Rockwell B hardness plotted against time at room temperature after
solutionizing and quenching into water (blue) and a polymer quenchent (green).



Table 4-1: Natural aging data from Figure 4-10 and Figure 4-11.
W-Temper T4-Temper Delta (W to T4)

HRB EC HRB EC Time (hr) HRB EC
Water Quench 60 30.1 76 28.6 ~12.6 16 -1.6
Polymer Quench 62 29.9 74 28.3 ~29.7 12 -1.6



113
28.0
28.5
29.0
29.5
30.0
30.5
0 10 20 30 40 50 60 70 80 9
Time (hr)
C
o
n
d
u
c
t
i
v
i
t
y

(
%
I
A
C
S
)
0
Water Quench
Polymer Quench
T4 (Water Quench)
W (Water Quench)
T4 (Polymer Quench)
W (Polymer Quench)

Figure 4-11: Eddy current electrical conductivity plotted against time at room
temperature after solutionizing and quenching into water (blue) and a polymer quenchent
(green).



The as-quenched (W-temper) hardness of the water quenched material is lower
than that of the polymer quenched. However, the hardness upon completion of the natural
aging process is highest for the water quenched material. The electrical conductivity of
the water quenched material is greater than the polymer in both the as-quenched and
stabilized condition while the magnitude of the drop is equivalent. The final discrepancy
between the water quenched and polymer quenched material is the time to reach the T4
condition. The polymer quenched material requires over twice the amount of time to
stabilize.
In an effort to find an optimum quenching rate that would be rapid enough to
prevent the formation of grain boundary precipitates while still avoiding mid-plane

114
cracking during rolling, the polymer quenchent was diluted. Different volume fractions of
polymer solution and water were mixed together for the material to be quenched into
after solutionizing. The as-quenched hardness, conductivity, amount of cold rolling strain
that could be imposed, and the estimated cooling rate were measured. This data is
graphically displayed in Figure 4-12 Figure 4-13 and . The specimens that were cold
rolled to strains exceeding 1.2 did not exhibit cracking and the rolling was stopped.


0
0.2
0.4
0.6
0.8
1
1.2
1.4
0 10 20 30 40 50 60 70 80 90 100
% Polymer Quenchent
T
r
u
e

R
o
l
l
i
n
g

S
t
r
a
i
n
0
5
10
15
20
25
E
s
t
i
m
a
t
e
d

T
i
m
e

t
o

R
e
a
c
h

1
0
0

C

(
m
i
n
)
Rolling Strain
Time

Figure 4-12: Maximum rolling strain applied before mid-plane cracking (material CR to
strains exceeding 1.2 did not crack) and estimated time to reach 100C plotted against the
percent polymer mixed into room temperature water.



115
29.6
29.7
29.8
29.9
30
E
d
d
y

C
u
r
r
e
n
t

C
o
n
d
u
c
t
i
v
i
t
y

(
%
I
A
C
S
)
60
61
62
63
64
65
H
a
r
d
n
e
s
s

(
R
o
c
k
w
e
l
l

B
)
HRB E.C.
0 10 20 30 40 50 60 70 80 90 100
% Polymer Quenchant

Figure 4-13: Rockwell B Hardness and Eddy Current Conductivity plotted against the
percent polymer mixed into room temperature water.


The 20/80 (polymer/water) mixture provided the most rapid quenching that also
allowed the material to be subsequently rolled without cracking, as shown in Figure
4-12. Note that the Time to reach 100C data read off of the secondary Y-axis are only
estimates and were not determined using any instrumentation. The hardness value at a
ratio of 20/80 corresponds to approximately 16% completion of the natural aging process.
Increasing the polymer ratio above 20% does not appreciably increase the hardness any
further. The conductivity data initially decreases as the quench rate is reduced, as shown
in Figure 4-13. At the 20/80 ratio this decreasing trend plateaus and then starts to
increase again around ~63/37.


116
4.4.1.2 Origin of Mid-Plane Cracking

It was hypothesized that the mid-plane cracking during rolling was a result of
either: (1) a high residual stress produced by the temperature gradient through the
coupons during the quenching process or (2) quench-cracks which are being propagated
by the rolling process rather than formed during it. Both of these working hypotheses are
directly related to the quench rate. What allowed for them to be isolated from each other
is that one is reversible (residual stress) while the other is not (quench-cracks). The
experimental procedure depicted in Figure 4-14 was the method used to isolate the two
possible causes of the cracking.



Figure 4-14: Flow chart depicting the experimental procedure used to isolate the two
possible causes of the cracking during cold rolling.
Water
Quench
Cold Rolling
Mid-Plane
Cracking
Polymer
Quench
Cold Rolling

No Cracking
Polymer
Quench
Solution
Treat
Cold Rolling

No Cracking
Water
Quench
Solution
Treat

117
It was already determined that quenching into water produced cracking during
subsequent rolling while quenching into non-diluted polymer quenchent did not. When
the material was SHT and water quenched followed by a second SHT and polymer
quench it was capable of being rolled to strains equivalent to previous polymer
quenching, as shown in Figure 4-15.



Figure 4-15: AA2524 plate that was successfully cold rolled after being solution heat
treated and water quenched followed by an additional solutionizing and polymer quench.


4.4.1.3 Uphill Quenching

The uphill quenching procedure was carried out as described in the Experimental
Methods section. The material was able to be successfully cold rolled in this condition
with no evidence of mid-plane cracking or distortion.

4.4.2 Deformation Processing - Cold Rolling

4.4.2.1 Hardness and Conductivity

The Rockwell B hardness and eddy current electrical conductivity profiles were
collected for the material cold rolled in the W-temper, half hardened condition (1/2H),
and the T4-temper. The profiles for the samples rolled in the half hardened condition are
shown in Figure 4-16 and Figure 4-17. The open data points represent the natural aging
progression from the as-quenched W-temper to the fully naturally aged T4-temper. The

118
solid data points represent samples that were cold rolled once the material had reached
half the T4 hardening increment (the hardening increment is the increase in hardness
relative to the as-quenched material). These profiles clearly show that imposing a plastic
strain alters the natural aging process. After 1.3 hours into the natural aging process the
samples hardness should increase an additional 7 HRB points and its conductivity should
drop an addition 0.9 %IACS as the formation of GP zones goes to completion. This
would occur gradually over the subsequent ~4 hours as it does for the open data points.
The gradual changes in hardness and conductivity are not what is observed for the
samples that were rolled. With the exception of the lowest rolling strain (
CR
= 0.15) the
hardness instantly increases upon cold rolling and remains constant. Any change in
hardness after deformation is within the noise of the measurement technique (1 hardness
point) and not believed to be a result of additional GP zone formation. The sample rolled
to a strain of 0.15 exhibits markedly different behavior. Upon rolling, the hardness of the
material increases slightly as a result of the introduction of newly generated dislocations,
however, over the next ~3.5 hour the hardness continues to increase the additional 7
points. This discrepancy in the behavior of material rolled to a strain below 0.2 and that
rolled to a strain at/or above 0.2 has previously been observed by the authors [125,
Appendix B]. It appears that there is a minimum amount of strain that must be imposed to
inhibit the natural aging process.



119
30
40
50
60
70
80
90
0 10 20 30 40 50 6
Time (hr)
H
a
r
d
n
e
s
s

(
H
R
B
)
0
Strain = 0
Strain = 0.15
Strain = 0.21
Strain = 0.41
Time of Rolling (1.3hr)

Figure 4-16: Rockwell B Hardness plotted against natural aging time (time at room
temperature after quenching) for the samples cold rolled in the half hardened condition
(1.3hr).



28.5
29.0
29.5
30.0
30.5
31.0
0 10 20 30 40 50 6
Time (hr)
E
l
e
c
t
r
i
c
a
l

C
o
n
d
u
c
t
i
v
i
t
y

(
%
I
A
C
S
)
0
Strain = 0
Strain =0.15
Strain = 0.21
Strain =0.41
Time of Rolling (1.3hr)

Figure 4-17: Eddy Current Electrical Conductivity plotted against natural aging time
(time at room temperature after quenching) for the samples cold rolled in the half
hardened condition (1.3hr).



120
The conductivity profiles corresponding to each of the hardness profiles in Figure
4-16 are shown in Figure 4-17. Upon cold rolling there is a sudden drop in the
conductivity. Following the initial drop, the conductivity increases until stabilizing. A
slight indication of a further conductivity drop is exhibited in the sample rolled to a strain
of 0.15. After the initial drop in conductivity there is a perturbation in the increasing
trend where the conductivity drops as the hardness is increasing. Similar trends were
observed in the Hardness/Conductivity profiles for the material deformed in the W-
temper and T4-temper.
The material deformed in the T4-temper exhibited the highest hardness values at
equivalent strains, as shown in Figure 4-18. Each data set was fitted to a linear trend line
and the slopes decrease as GP zones form. The slope of the W-temper data set is almost
2.5 times larger than that of the T4-temper data set. The relevance of this observation is
that while the samples deformed in the T4-temper exhibit a higher final hardness, a
greater increment of hardness was achieved by the samples deformed in the W-temper, as
shown in Figure 4-19. The hardening increment is the increases in hardness resulting
only form the increases in dislocation density. The increased hardness resulting from
natural aging before rolling and the additional hardening observed after rolling for the W
and 1/2H samples rolled to strains below 0.2 was subtracted out.

121
y = 62.138x + 60.08
y = 36.404x + 71.503
y = 26.33x + 76.258
0
10
20
30
40
50
60
70
80
90
100
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45
True Rolling Strain
T
o
t
a
l

H
a
r
d
n
e
s
s

(
H
R
B
)
W-Temper
1/2H-Temper
T4-Temper

Figure 4-18: Rockwell B Hardness plotted against Rolling Strain with the data for each
temper fitted to a linear trend line.




122
y = 23.696x + 1.2705
y = 34.95x + 1.9776
y = 49.366x + 7.5485
0
5
10
15
20
25
30
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45
True Rolling Strain
H
a
r
d
n
e
s
s

I
n
c
r
e
m
e
n
t

(
H
R
B
)
W-Temper
1/2H-Temper
T4-Temper

Figure 4-19: The Hardness Increment resulting from dislocations only plotted against
Rolling Strain with data for each temper fitted to a linear trend line.


The drop in electrical conductivity upon rolling follows a similar trend as that
observed for the hardening increments for the different tempers. In the absence of GP
zones not only is the magnitude of the drop larger at equivalent strains, but the rate of
change is also considerably larger than in the presence of GP zones, as shown by the
slope of each data set in Figure 4-20. There is a positive correlation between
conductivity drop and increased hardening that becomes clearer if the values are plotted
against each other, as shown in Figure 4-21. Note that the Y-intercept does not pass
through the origin although the intercept value is within the range of the noise of the
hardness measurements.

123
y = 0.6204x + 0.1606
y = 1.3004x + 0.4762
y = 2.0094x + 0.41
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45
Rolling Strain
C
o
n
d
u
c
t
i
v
i
t
y

D
r
o
p

(
%
I
A
C
S
)
W-Temper
1/2H-Temper
T4-Temper

Figure 4-20: Eddy Current Electrical Conductivity Drop upon rolling plotted against
rolling strain.




124
y = 24.343x - 2.31
y = 29.732x - 13.134
y = 33.791x - 3.4321
y = 20.157x - 0.4938
0
5
10
15
20
25
30
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
Conductivity Drop (%IACS)
H
a
r
d
n
e
s
s

I
n
c
r
e
m
e
n
t

(
H
R
B
)

W-Temper
1/2H-Temper
T4-Temper
Linear (All Data)

Figure 4-21: The Hardness Increment resulting from dislocations only plotted against
eddy current electrical conductivity drop upon rolling.


4.4.2.2 Strengthening and SHR Data

Yield strength data was collected form the material used to produce the plots in
Figure 4-18 through Figure 4-21. The compressive yield strengths were plotted against
the corresponding rolling strains and fitted to a power law in order to take the form of the
Hollomon equation [126]:
(4.4)
n
Kc o =
where K is the strain hardening parameter and n is the strain hardening exponent, as
shown in Figure 4-22. When the data sets are fitted to the relationship, it is seen that both
the strain hardening coefficient (K) and the strain hardening exponent (n) decrease as GP
zones form, as shown in Table 4-2. What these parameters predict is that even though the

125
material in the W-temper starts at a lower yield strength it will reach the same flow stress
as the T4 material at a strain of approximately 0.45 and stress of 487MPa. The
compressive stress-strain curves for each of the fourteen combinations of aging and
rolling strain are in Appendix A.


Table 4-2: Data from Figure 4-22 in tabular form.
W-Temper (Ave.) H-Temper (Ave.) T4-Temper (Ave.)
CR
Strain
Yield
Strength
(MPa)
Yield
Strength
(Ksi)

CR
Strain
Yield
Strength
(MPa)
Yield
Strength
(Ksi)

CR
Strain
Yield
Strength
(MPa)
Yield
Strength
(Ksi)
0.00 148 21.5 0.00 221 32.1 0.00 321 46.5
0.13 393 57.0 0.15 434 63.0 0.12 424 61.5
0.23 431 62.5 0.21 455 66.0 0.22 465 67.5
0.31 452 65.5 --- --- --- 0.31 479 69.5
0.41 469 68.0 0.41 486 70.5 0.41 496 72.0
K 570 MPa K 555 MPa K 515 MPa
n 0.91 n 0.13 n 0.07



126
y = 569.55x
0.1941
R
2
= 0.9988
y = 554.78x
0.1327
R
2
= 0.9992
y = 515.38x
0.07
R
2
= 0.9761
0
100
200
300
400
500
600
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45
True Cold Rolling Strain
C
o
m
p
r
e
s
s
i
v
e

Y
i
e
l
d

S
t
r
e
n
g
t
h

(
M
P
a
)
0
10
20
30
40
50
60
70
80
C
o
m
p
r
e
s
s
i
v
e

Y
i
e
l
d

S
t
r
e
n
g
t
h

(
K
s
i
)
W-Temper (Ave.)
H-Temper (Ave.)
T4-Temper (Ave.)

y
= K
n

Figure 4-22: Compressive yield strength plotted against cold rolling strain for the each of
the three tempers.


The Kocks [127] plot is a graphical representation of how the SHR progresses per
increment of stress above the yield strength, as shown in Figure 4-23. Representing the
SHR data in this manner is advantageous because it allows for the effect of only
dislocation density on strengthening to be analyzed. The Kocks plot has the effect of
superimposing the yield points of each data set on top of each other, eliminating the
contribution of other strengthening mechanisms.

127
0
500
1000
1500
2000
2500
3000
3500
4000
4500
5000
0 50 100 150 200 250 300 350
CR Flow - Yield Stress (MPa)
S
H
R

(
M
P
a
)
0
100
200
300
400
500
600
700
0 10 20 30 40 50
CR Flow - Yield Stress (ksi)
S
H
R

(
k
s
i
)
W-Temper
W-Temper (equ.)
1/2H-Temper
1/2H-Temper (equ.)
T4-Temper
T4-Temper (equ.)
Series7
Power (W-Temper)
Power (1/2H-Temper)
Power (T4-Temper)

Figure 4-23: Strain Hardening Rate (SHR) plotted against the Cold Rolling (CR) flow
stress minus the Yield Stress.


The solid data points represent the numerical SHR data determined using the
values in Table 4-2, the broken lines represent a power law fit of the numerical data
points, and the continuous solid lines are the SHRs determined using the Hollomon
relationship in Figure 4-22. The SHR data determined using the Hollomon relationship is
believed to be the most accurate representation of the material's behavior. The trend lines
fitted to the numerical data underestimate the SHR. This is a result of the relatively large
spacing between the stress-strain points in Figure 4-22. It is clear that at an equivalent
flow stress the SHR is higher in the absence of GP zones regardless of which set of
curves are compared. This directly confirms that the SHR of AA2524 is more potent
before GP zone formation.

128
If the compressive yield strength increment (the increment of strength from only
strain hardening with the effects of aging subtracted out) is plotted against the
conductivity drop upon rolling (the increase in resistivity is believed to scale with
dislocation density) a square root relationship is exhibited, as shown in Figure 4-24. This
is an interesting observation being that shear stress and flow stress scale linearly and it is
believed that the drop in conductivity scales with the dislocation density.

y = 283.64x
0.5349
R
2
= 0.9522
0
50
100
150
200
250
300
350
0.00 0.20 0.40 0.60 0.80 1.00 1.20 1.40
Final Conductivity Drop (%IACS)
C
o
m
p
r
e
s
s
i
v
e

Y
i
e
l
d

S
t
r
e
n
g
t
h

I
n
c
r
e
m
e
n
t

(
M
P
a
)
0
10
20
30
40
50
C
o
m
p
r
e
s
s
i
v
e

Y
i
e
l
d

S
t
r
e
n
g
t
h

I
n
c
r
e
m
e
n
t

(
k
s
i
)
W-Temper
1/2H-Temper
T4-Temper
Power (All Data)

Figure 4-24: Compressive yield strength increment plotted against the conductivity drop
upon rolling.








129
4.4.3 Compression Testing Data

Compressive stress-strain curves were constructed at each level of cold rolling
strain. The microstructures of the specimen tested at a strain of zero, shown in Figure
4-25, only differ in the amount of GP zones present. The continuous SHR curves
generated from the compression tests for the strain free material are in agreement with the
discrete SHR curves, Figure 4-23, for the rolled material. As GP zones form, the yield
strength increases and the SHR decreases as shown in Figure 4-25, Figure 4-26, and
Figure 4-27.

0
100
200
300
400
500
600
700
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
True Compressive Strain
T
r
u
e

C
o
m
p
r
e
s
s
i
v
e

S
t
r
e
s
s

(
M
P
a
)
0
20
40
60
80
100
T
r
u
e

C
o
m
p
r
e
s
s
i
v
e

S
t
r
e
s
s

(
k
s
i
)
W-Temper
1/2H-Temper
T4-Temper
y = 783.28x0.3289
R2 = 0.9975
y = 768.08x0.2544
R2 = 0.9977
y = 816.23x0.2024
R2 = 0.9959

Figure 4-25: Compression stress-strain curves for the non-rolled material tested in the W,
1/2H, and T4 temper.


130
0
500
1000
1500
2000
2500
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40
True Compressive Strain
S
H
R

(
M
P
a
)
0
50
100
150
200
250
300
350
S
H
R

(
k
s
i
)
W-Temper
1/2H-Temper
T4-Temper

Figure 4-26: SHR plotted against compressive strain for the non-rolled material tested in
the W, 1/2H, and T4 temper.




131
0
1000
2000
3000
4000
5000
6000
7000
8000
0 50 100 150 200 250 300 350 400
True Compressive Flow - Yeild Stress (MPa)
S
H
R

(
M
P
a
)
0
200
400
600
800
1000
0 10 20 30 40 50
True Compressive Flow - Yeild Stress (ksi)
S
H
R

(
k
s
i
)
W-Temper
1/2H-Temper
T4-Temper


Figure 4-27: SHR plotted against flow stress with the yield stress subtracted for the non-
rolled material tested in the W, 1/2H, and T4 temper.



The material tested in compression after rolling was subjected to a strain path
change unlike the material tested at zero strain. The change in strain path has a temporary
but significant effect on the strain hardening response regardless of the temper, as shown
in Figure 4-28 through Figure 4-33. At the onset of compression, the SHR for each
temper is initially higher for the more heavily rolled material. However, at higher rolling
strains the SRH diminishes the most rapidly until it coincides with non-rolled material.
This transient material response is seen when the SHR is plotted against total strain
(rolling strain plus compressive strain), as show in Figure 4-28, Figure 4-30, and Figure
4-32. In the absence of a strain path change, it would be expected that the data sets would
lie directly on top of each other. Note that the power law relationship that the data was

132
fitted to was only able to capture the SHR response once the initially very high rate had
diminished. Therefore the initial, almost vertical, segment of each data set is not
displayed. This transient SHR effect resulting from the change in strain path was
magnified by GP zone formation.
The Kockes plots, Figure 4-29, Figure 4-31, and Figure 4-33, reveal this
transient SHR effect as an intersection of the data sets. It is initially seen that the SHR
increases with increased prior rolling (counter intuitive if there had not been a strain path
change) but as compression continues the lower rolling strains exhibit higher SHR. These
data sets will not coincide until State IV hardening is reached in all samples.






133
0
500
1000
1500
2000
2500
3000
3500
4000
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Total (CR+Compression) True Strain
S
H
R

(
M
P
a
)
0
100
200
300
400
500
S
H
R

(
k
s
i
)
Log. (T4_E=0-All)
Power (T4_E=0.12-All)
Power (T4_E=0.22-All)
Power (T4_E=0.31-All)
Power (T4_E=0.41-All)

Figure 4-28: SHR plotted against total strain (rolling strain plus compressive strain) for
the material cold rolled in the T4-temper.
0
1000
2000
3000
4000
5000
6000
7000
0 50 100 150 200 250 300 350
True Compressive Flow - Yield Stress (MPa)
S
H
R

(
M
P
a
)
0
200
400
600
800
1000
0 10 20 30 40 50
True Compressive Flow - Yield Stress (kis)
S
H
R

(
k
s
i
)
Log. (T4_E=0-All)
Poly. (T4_E=0.12-All)
Poly. (T4_E=0.22-All)
Poly. (T4_E=0.31-All)
Poly. (T4_E=0.41-All)

Figure 4-29: SHR plotted against flow stress with the yield stress subtracted for the
material cold rolled in the T4-temper.

134
0
500
1000
1500
2000
2500
3000
3500
4000
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Total (CR + Compression) True Strain
S
H
R

(
M
P
a
)
0
100
200
300
400
500
S
H
R

(
k
s
i
)
Log. (1/2H_E=0.0-All)
Power (1/2H_E=0.15-All)
Power (1/2H_E=0.21-All)
Power (1/2H_E=0.41-All)

Figure 4-30: SHR plotted against total strain (rolling strain plus compressive strain) for
the material cold rolled in the 1/2H-temper.
0
1000
2000
3000
4000
5000
6000
7000
8000
9000
10000
0 50 100 150 200 250 300 350 400
True Compressive Flow - Yield Stress (ksi)
S
H
R

(
M
P
a
)
0
200
400
600
800
1000
1200
1400
0 10 20 30 40 50
True Compressive Flow - Yield Stress (ksi)
S
H
R

(
k
s
i
)
Log. (1/2H_E=0.0-All)
Poly. (1/2H_E=0.15-All)
Poly. (1/2H_E=0.21-All)
Poly. (1/2H_E=0.41-All)

Figure 4-31: SHR plotted against flow stress with the yield stress subtracted for the
material cold rolled in the 1/2H-temper.

135
0
500
1000
1500
2000
2500
3000
3500
4000
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Total (CR + Compression) True Strain
S
H
R

(
M
P
a
)
0
100
200
300
400
500
S
H
R

(
k
s
i
)
Log. (W_E=0.0-All)
Power (W_E=0.41-All)
Power (W_E=0.13-All)
Power (W_E=0.23-All)
Power (W_E=0.31-All)

Figure 4-32: SHR plotted against total strain (rolling strain plus compressive strain) for
the material cold rolled in the W-temper.
0
1000
2000
3000
4000
5000
6000
7000
8000
0 50 100 150 200 250 300 350 400 450
Compressive Flow - Yield Stress (MPa)
S
H
R

(
M
P
a
)
0
200
400
600
800
1000
0 10 20 30 40 50 60
Compressive Flow - Yield Stress (ksi)
S
H
R

(
k
s
i
)
Log. (W_E=0.0-All)
Poly. (W_E=0.13-All)
Poly. (W_E=0.41-All)
Poly. (W_E=0.31-All)
Poly. (W_E=0.23-All)

Figure 4-33: SHR plotted against flow stress with the yield stress subtracted for the
material cold rolled in the W-temper.

136
4.5 Discussion

4.5.1 Heat Treating

In this study the solution heat treatment of the alloy is arguably the most critical
step of the processing. The objective of this operation is to produce the highest level of
super saturation in solid solution feasible. This is achieved by solutionizing at the
maximum temperature below the eutectic melting point followed by cooling as rapidly as
possible. The rapid cooling rate locks the solute in solution preventing its loss to the
formation of lower energy phases.

4.5.1.1 Quench Rate

The cracking observed during rolling was not anticipated being that there was no
mention of any such problems in the work that established that 2xxx series alloys could
be effectively strain hardened [5,6], nor were there any such problems mentioned in other
relevant studies [3,4]. The inability to deform the material had to be overcome in order to
proceed with the investigation. It was believed that there were two possible sources that
could be producing the mid-plane cracking. The first was the non-uniform cooling within
the sample producing a sustained stress in the material after quenching. The temperature
gradient created through the thickness of the sample during quenching most likely would
not have been an issue in the work conducted by [5,6]. The reason for this is that the
work was carried out on 3.2mm thick sheet. At this gauge the heat would be rapidly
drawn out of the material and the temperature differential between the surface and the
center of the sheet may not have been large enough to produce a residual stress of

137
appreciable magnitude. The material in the current study is more than three times as thick
(10mm). The current work is part of a larger study to determine the feasibility of using
solute enhanced strain hardening (SESH) to achieve better combinations of strength and
toughness in Al alloys for load bearing applications. The nature of this work (and the
ultimate application) requires a thicker gauge material and it is known that as the
thickness of the material is increased, so is the residual stress produced by quenching
[117, 122]. The second possible source of the mid-plane cracking may have been cracks
produced during the rapid expansion and contraction during quenching. That is, that the
cracks seen during rolling may have been preexisting quench-cracks that were propagated
by the rolling process rather then created by it. The two possible sources (residual stress
or quench cracks) of the cracking are both directly dependent on the quench rate.
It was expected that reducing the quench rate would affect the resulting
microstructure and the natural aging process of alloy 2524. Although it was not known if
the resulting material would still be suitable for the overall investigation. It has long been
established that GP zone formation is accompanied by an increase in hardness and
decrease in electrical conductivity; both of which plateau when the natural aging process
has gone to completion, producing material in the well recognized T4-temper. The
hardness and conductivity profiles for the two quenching rates (water and polymer) are
shown in Figure 4-10, Figure 4-11, and Table 4-1.
The discrepancies between the natural aging profiles of the water and polymer
quenched material is a result of the as-quenched microstructures. The water quenched
material (at time equal to zero) has an aluminum matrix super saturated with solute. The
polymer quenched material also has an aluminum matrix super saturated with solute,

138
although some of the solute has been lost to both GP zones and non-coherent grain
boundary precipitates. It is also expected that the slower quench rate will result in a lower
volume fraction of excess vacancies.
This loss of solute lowers the level of super saturation and the driving force to
form solute clusters. The slower quench rate also results in a lower volume fraction of
excess vacancies which fuel the diffusion process that allows for the clustering of solute
atoms at room temperature.
The higher hardness observed for the polymer quenched material is a result of GP
zone formation during the quenching process. This loss of solute to GP zones is reflected
in the conductivity profile as a lower as-quenched value, relative to the water quenched
material. The lower final hardness of the polymer quenched material is a result of grain
boundary precipitation during the quench. Solute lost to these precipitates has no
contribution to hardness (strengthening) and is unavailable to cluster into strengthening
GP zones. The most substantial difference between the materials quenched at different
rates is the time required to reach the T4 condition. This is due to the polymer quenched
material having a lower level of super saturation and quenched-in vacancies. The lower
level of excess vacancies has a profound effect on the rate of diffusion. The diffusion
coefficients for the two quench rates was estimated using the relationship:

t
D
4
2

=
(4.4)
and the diffusion coefficient for Cu in Al was extrapolated down to room temperature
using the relationship:

RT
Q
o
e D D

=
(4.5)

139
where is the GP zone spacing, t is time to stabilize, T is temperature, R is the gas
constant, D
o
is the experimentally determined frequency factor, and Q is the
experimentally determined activation energy of diffusion. The estimated values are given
in Table 4-3 and Figure 4-34. It was assumed that the GP zone spacing in the water
quenched and polymer quenched material are equivalent. This is believed to be a
reasonable assumption being that the magnitude of the conductivity drops during natural
aging were equivalent, as shown in Figure 4-11.

Table 4-3: Values and variables used in equations 4.4 and 4.5 to construct Figure 4-34
[128].
Condition
D
(m^2/sec)
Do
(m^2/sec)
Q
(KJ/mol)
Time
(hr)
Temperature
( C )
Spacing
(m)
Published Data -- 6.54E-05 136 -- 321 - 655 --
Extrapolated Data -- 6.54E-05 136 -- 0 - 640 --
Water Quenched 5.5E-22 -- -- 12.6 20 1.E-08
Polymer Quenched 2.3E-22 -- -- 29.7 20 1.E-08
Room Temp. 3.8E-29 6.54E-05 136 20,848 yr 20 --





140
5.50E-22
3.82E-29
2.34E-22
1.E-30
1.E-28
1.E-26
1.E-24
1.E-22
1.E-20
1.E-18
1.E-16
1.E-14
1.E-12
1.E-10
0 100 200 300 400 500 600 700
Temperature (C )
D

(
m
2
/
s
e
c
) Extropolated Data
Published Data
Water Quenched
Polymer Quenched
Room Temperture

Figure 4-34: Diffusion Coefficient plotted against temperature for published high
temperate data extrapolated down to temperatures that natural aging occurs at and the
experimentally estimated values for different quenching rates.



The diffusion coefficient of the polymer quenched material is ~58% lower than
that of the water quenched material. The diffusion coefficient at room temperature in the
absence of excess vacancies is orders of magnitude lower and, if the time for natural
aging to go to completion is backed out of equation 4.4, it predicts a period of over
20,000 years.
The diffusion coefficient of the water quenched material being orders of
magnitude larger then that predicted for Cu in Al at room temperature is supported by the
estimated concentrations of excess vacancies predicted in each condition. The
equilibrium concentration of vacancies at a given temperature can be determined using
the relationship:

141

KT
E
v
vf
e T C

= ) (
(4.6)
where E
vf
[129] is the energy of formation and K is Boltzmanns constant. When the
material is rapidly quenched form the SHT temperature the equilibrium concentration of
vacancies at the elevated temperature is locked in at room temperature (in the same
manner that the solute is locked in). The magnitude of excess vacancies immediately after
quenching relative to the equilibrium concentration is shown in Figure 4-35. In the as-
quenched condition there are approximately 3x10
8
vacancies for every one in equilibrium
at room temperature. These excess vacancies act to transport solute [96] and form GP
zones. As the excess vacancies are annealed out at room temperature the natural aging
process will slow until it ultimately halts. It should be made clear that it is not the natural
aging process that eliminates the excess vacancies, but rather the elimination of the
excess vacancies that stops the natural aging process. The excess vacancies are
eliminated by sinks such as grain boundaries, phase boundaries, free surfaces, interfaces,
and dislocations [93].

142
1.E-15
1.E-12
1.E-09
1.E-06
1.E-03
0 100 200 300 400 500 600 700
Temperture (C)
V
a
c
a
n
c
y

C
o
n
c
e
n
t
r
a
t
i
o
n

Equilibrium Concentration
Concentration at SHT Temp.
Water Quenched
Room Temp.
Water Quenching
N
a
t
u
r
a
l

A
g
i
n
g

Figure 4-35: Vacancy concentration plotted against temperature showing the excess
quenched-in vacancies present after SHT.



The vacancy concentration present immediately after polymer quenching was not
able to be estimated with any level of certainty due to solute being lost to grain boundary
precipitates during the quench.
While the formation of GP zones during quenching is undesirable it is an
acceptable alternative to mid-plane cracking during rolling. The formation of grain
boundary precipitates during the quench, however, is not an acceptable alternative. The
polymer quenchent was diluted in an attempt to produce a quench rate that would be
rapid enough to prevent grain boundary precipitation while also avoiding mid-plane
cracking during cold rolling. The rolling, quench rate, hardness, and conductivity results

143
for various volume fractions of polymer quench are shown in Figure 4-12 and Figure
4-13.
It was determined that a 20/80 (polymer/water) mixture provided the most rapid
quenching which also allowed the material to be subsequently rolled without cracking.
The relevant question is whether or not a 20/80 ratio provides a rapid enough quench to
prevent the formation of embrittling grain boundary precipitates. As would be expected,
the as-quenched hardness of the material increases as the quench rate is reduced due to
the formation of GP zones. The increase in hardness with decreasing quench rate was
reasonably well fit to a power law relationship which captured the initial rapid increase.
The electrical conductivity data, read off the secondary Y-axis in Figure 4-13, provides
greater insight into any phase transformations that may be occurring during the quench.
The conductivity data was fit to a third order polynomial. It is not being suggested that
either the change in hardness or conductivity with decreasing quench rate is accurately
described by the functions used to fit the data. What is being claimed is that the trend
lines capture the overall behavior of the data and are useful visualization tools for
analysis. The parabolic shape of the conductivity data is the result of two concurrent
phase transformations. The initial drop in conductivity is characteristic of GP zone
formation. The clustering of solute atoms (GP zones) scatters electrons more effectively
then isolated solute atoms. As the quench rate is further reduced, the conductivity starts to
increase. An increasing trend in conductivity is observed during the artificial aging
process where the matrix becomes more solute free as precipitates form. There is not a
large enough driving force available during the quenching process to overcome the
barrier to nucleation for the coherent and semi-coherent strengthening precipitates to

144
form. Therefore, the observed change in the decreasing trend for the electrical
conductivity data at a ratio of 20/80 is a result of the net conductivity change resulting
from GP zone formation (decreasing conductivity) and the formation of grain boundary
precipitates (increasing conductivity).
The data collected for the various cooling rates clearly shows that the quench
directly impacts the formation of embrittling grain boundary precipitates and the ability
of the material to be rolled. What was unable to be determined with an adequate level of
confidence was a polymer/water quenchent ratio that would reproducibly provide a
cooling rate that would both facilitate rolling and prevent the formation of grain boundary
precipitates. Form the data shown in Figure 4-12, it is known that a slower cooling rate
than what was produced by the 10/90 solution is required to avoid mid-plan cracking,
while the data in Figure 4-13 shows that a more rapid cooling rate then what is produced
by a 20/80 solution is necessary to avoid the formation of grain boundary precipitates
during the quench. Searching for the optimum ratio within the 10-20 vol% polymer
solution range was not further explored due to the practicality of operating within such a
tight window. This is in part due to the fact that the synthetic quenchent degrades over
time and with repeated use. The other major factor in deciding not to further explore
quenching within the 10-20 vol% was the severity of the consequences associated with
slightly deviating outside of the acceptable window (cracking and grain boundary
precipitation).




145
4.5.1.2 Origin of Mid-Plane Cracking

The decision to abandon using a synthetic quenchent to avoid the mid-plane
cracking phenomena required the author to revisit the question of what the root cause of
the cracking was. It was hypothesized earlier in the discussion that the mid-plane
cracking during rolling was a result of either: (1) a high residual stress produced by the
temperature gradient through the coupons during the quenching process or (2) quench-
cracks which are being propagated by the rolling process rather than formed during it.
While the polymer quenchent did not provide a solution to the cracking problem, it would
aid in determining the root-cause of the cracking. Both of the working hypotheses are
directly related to the quench rate. What allowed for them to be isolated from each other
is that one is reversible (residual stress) while the other is not (quench-cracks). The
experimental procedure depicted in Figure 4-14 was the method used to isolate the two
possible causes of the cracking.
It was already determined that quenching into water produced cracking during
subsequent rolling while quenching into non-diluted polymer quenchent did not. What
was not know was if the synthetic quenchent was preventing cracking by reducing the
temperature gradient through the thickness of the material resulting in a lower residual
stress, or if it was reducing the rate of thermal contraction/expansion preventing quench-
cracks. Being that quench-cracks are exposed to the environment, re-solution heat
treating and cold rolling will not heal them (like HIPing would heal hydrogen casting
porosity). As a result, if a coupon was solutionized and water quenched, producing
quench-cracks, followed by a second solution heat treatment and polymer quench, the
coupon would still crack upon rolling. When the experiment was conducted, the double

146
solutionized coupon did not crack but rather was able to be rolled to a strain of 2. This
indicates that the mid-plane cracking is a result of residual stress and not quench-cracks.
This conclusion is supported by the lack of observable (both at magnification and with
die penetrant inspection) cracks after water quenching, geometric distortion observed
during machining of the water quenched material, as shown in Figure 4-36, and the
ability to roll the material after performing the uphill quenching process.



Figure 4-36: Shape change (curling) of water quenched material during an EDM surface
pass. The normal to surface of the work piece faces toward the center of the curling
indicating a residual tensile surface stress.
15 mm


During the rolling process, the mid-plane cracking isnt observed until after ~3
passes through the rolling mill. It was initially thought that the strain being imposed
during these first few passes would mechanically relieve any residual stress present after
quenching. This was not, however, the case due to the non-uniformity of the imposed
strain inherent to the rolling process (unlike uniaxial tension or simple compression used
to mechanically stress relieve material).

4.5.1.3 Uphill Quenching

When heat-treatable Al alloys are commercially produced they are mechanically
stress relieved by imposing a uniform strain of 2-3%. This is typically done by stretching.

147
Stretching is an undesirable method for the present study being that the amount of
imposed strain is one of the primary variables, and stretching would introduce both
additional strain and an additional strain path (uniaxial tension vs. plane strain
compression). Another widely used method of stress relieving is thermal annealing. This
is also not a viable alternative for this experiment because phase transformations would
occur in the temperature range required to effectively anneal, resulting in the depletion of
the of solute from the matrix. At the extreme upper temperature range, the material would
be re-solutionized and the residual stress would once again be imparted on the material
during the required rapid cooling. For these, as well as other reasons, the uphill quenched
process was determined to be the best solution to the residual stress problem.
The uphill quenching process described in the Processing and Experimental
Method section is a thermal-mechanical process which exposes the material to a
temperature gradient in the opposite direction as the SHT quench. This imparts a stress of
equivalent state but opposite direction of the already existing sustained stress. The
magnitude of the stress imposed during the uphill quenching process relative to that
already in the material determines the effectiveness of the process. Typically this
thermal-mechanical stress relieving is carried out using liquid nitrogen and super heated
steam to produce a larger temperature differential during the uphill quench. In the present
study boiling water was used in place of high pressure steam due to the ease and
reproducibly of producing it.
Utilizing the uphill quenching process to thermal-mechanically stress relieve the
coupons allowed for the study to proceed without having to impose an additional strain or
embrittle the material by producing grain boundary precipitates.

148
4.5.2 Deformation Processing - Cold Rolling

In an earlier study conducted by the author (found in Appendix B, [125]) the
response of alloy AA2524 to a given amount of rolling strain as a function of natural
aging time was investigated. The hardness and conductivity data collected during the
study provided valuable insight into the response of the material to plastic deformation
during the natural aging process. The current investigation further expands on this
knowledge by directly measuring the yield strength and SHR over a larger range of
rolling reductions. The hardness and electrical conductivity trends, as a function of both
CR strain and natural aging time, were also reconfirmed with greater accuracy.

4.5.2.1 Hardness and Conductivity Trends

The material deformed in the T4-temper exhibited the highest hardness values at
equivalent strains, as shown in Figure 4-18. This material has the advantage of benefiting
from two separate strengthening methods, the resistance to slip provided by the GP zones
(soft shearable obstacles) and the increased dislocation density (a combination of hard
and soft obstacles) from the CR processing. Although the T4 material exhibits the overall
highest hardness values, the rate at which the hardness is increasing with increased strain
is the lowest. The lower hardening rate for the T4 data set is indicated by the slope of the
trend line. The lower hardening rate of the T4 material becomes more apparent when the
hardening increment resulting from only the imposed strain is plotted (the increment due
to the increase in dislocation density), as seen in Figure 4-19. The decrease in the
hardening increment at a given strain, as indicated by the decreasing slopes of the data

149
sets as GP zones form, strongly supports strain hardening being more potent when solute
is in solution rather then in clusters. It should be recognized that it is not expected that the
data in Figure 4-18 or Figure 4-19 should have a linear relationship. Linear
approximations are often used to correlate hardness measurements to yield strength [130],
and it is know that yield stress and rolling strain have a power law relationship. The
linear approximation in these plots was used to more clearly illustrate the rates of change
in the data sets relative to each other.
The drop in electrical conductivity upon rolling follows a similar trend as that
observed for the hardening increments of the different tempers, as shown in Figure 4-20.
The larger conductivity drops in the W-temper data reflect what is thought to be the
retention of a larger number of dislocations for a given imposed strain. The greater
retention of dislocations in the W material further supports an increased hardening rate in
the absence of GP zones. This positive correlation between dislocation density
(conductivity drop) and increased hardening is more clearly shown in Figure 4-21.
The hardness and conductivity data shows strong evidence that deforming
AA2524 in the absence of GP zones results in a higher retention of dislocations,
producing a larger increment of hardening and larger drop in conductivity than if
deformed in the T4-temper.
Hardness is not a material property, but rather an indicator of other material
properties. With units of stress, hardness is a function of both yield strength and SHR.
Before the conclusion can be made that AA2524 strain hardens more effectively with
solute in solution as opposed to solute clustered into GP zones, the yield strength and
SHR of each condition needs to be characterized.

150
4.5.2.2 Strengthening and SHR Data

The flow stress of the material deformed in the W-temper exhibits a higher SHR
than the material deformed containing GP zones, as shown by the strain hardening
exponents in Table 4-2. Dislocations are being retained at a higher rate in the presence of
solute atoms in solution as opposed to clustered into zones. The effect of the higher
retention of dislocations is clearly depicted in the Kocks' plot shown in Figure 4-27.
It was previously thought that SESH was a result of GP zones decreasing the SFE
of the alloy, hindering cross-slip. The formation of GP zones is no longer thought to
lower the SFE nor do the experimental observations of the current study support the SFE
theory. If the SFE of the alloy was being significantly lowered as the location of where
the solute resided changed from:
solution saturated Super zones GP ecipitates Pr

than it would be expected that deformation twins would form within the microstructure.
Deformation twins were not observed in the current work nor where they reported in any
of the previous SESH studies. Note that previous investigators predicted that GP zones
where the most effective microstructure feature at decreasing the SFE. However, the
super saturated solution was added to the end of the above sequence in light of the current
work and the manner in which solute atoms interact with stacking faults. The formation
of GP zones would be expected to increase the SFE relative to solute atoms in solution
not decrease it. The solute atoms often have a higher solubility in the disrupted crystal
structure of the stacking fault. The higher solubility results in the solute atoms
preferentially segregating to stacking faults, effectively lowering the SFE. The stability of

151
stacking faults would be reduced by solute clustering into GP zones and the SFE
increased.
Another possible explanation for the increased SHR in the W-temper is based on
the relative spacings of GP zones and dislocations. If the dislocation density rapidly
increased during deformation to a point where the mean dislocation spacing was less then
the mean GP zone spacing the strengthening effects of the zones would be secondary to
dislocation-dislocation interaction. In this situation, strain hardening is diminished
because the dislocation interactions are occurring in a higher purity matrix as a result of
the solute being lost to zone formation. As shown in the Figure 4-40, a dislocation
density of greater then 2.7E+16 would be required to produce a mean spacing of 10nm
(GP zone spacing) which would correspond to an increase in flow stress of over
1200MPa. Strain hardening would be exhausted long before this dislocation spacing was
reached. Therefore, the higher SHR in the W-temper cannot be explained by the GP zone
spacing diminishing favoring strengthening due to dislocation-dislocation interactions in
a solute depleted matrix.
The explanation of the higher SHR in the W-temper put forth in this study is that:
SESH is more effective in the presence of a super saturated matrix than one containing
GP zones a result of: (1) an increased frictional stress and (2) a lower rate of dynamic
recovery. The increase in frictional stress is thought to be a result of the interactions
between the long range stress fields of the dislocations with those of the short range stress
fields of the solute atoms. This is analogues to solution strengthening where a Cottrell
atmosphere is formed and a dislocation either has to overcome the attractive forces to
glide or drag the solute atom with it. Both of which will increase the resistance to flow. If

152
the solute atoms have clustered into GP zones they are no longer available to migrate to
dislocations. The dislocations then interact with the more widely spaced shearable GP
zones rather then the individual solute atoms. It is not known what effect the shearing of
the GP zones has on the SHR. The damaging of the GP zones may result in a more rapid
decline in the SHR of the T4 material. The experimental results suggest that the
interaction with GP zones is not as effective at inhibiting dislocation motion as solute
atoms in solution. The higher retention of dislocations (inhibited dynamic recovery) is
believed to come from solute atoms reducing the attraction of dislocations of opposite
signs to each other and inhibiting cross-slip by reducing the ability of a dislocation to
change slip planes. Dislocation-vacancy interaction may also increase dislocation line
length. Detailed TEM studies are required to confirm these mechanisms.
The higher retention of dislocations in an Al matrix containing solute in solution
rather then clustered into GP zones is further confirmed by Figure 4-24. It is known that
the shear stress of a metal scales with square-root of the dislocation density. If the
compressive yield strength is plotted against the conductivity drop upon rolling (the
conductivity drop is believed to scale with dislocation density) this same square root
relationship is exhibited. This observation supports both the hypothesis that the origin of
the conductivity drop is a result of an increased dislocation density and that solute in
solution impedes dynamic recovery more effectively than GP zones.





153
4.5.3 Compression Testing Data

In the previous section the SHR of AA2524 as a function of GP zone formation
and imposed CR strain was discussed. The current discussion will focus on the SHR at
each increment of CR strain. This type of testing provides insight into a materials
capacity to further deform and/or absorb energy after processing.
The continuous stress-strain and SHR data generated by compression testing at
zero-time is in agreement with the discrete cold rolling data. The compression data
showed that as GP zone formation occurs, the yield strength increases (as expected) and
the SHR decreases, as shown in Figure 4-28 through Figure 4-33. While the general
trend of the data for the different deformation modes are in agreement, the material tested
in compression shows a higher flow stress and SHR then the rolled material at equivalent
strains. The strain hardening exponent (n) for the CR material is 40% lower then the
material tested in compression when in the W-temper and this difference increases as GP
zones form, as shown in Figure 4-37. The difference in the strain hardening parameter
(K) also increases at longer aging times, a shown in Figure 4-38. The increased K value
in the T4-temper is a result of the higher yield stress and relative differences in n values.

154
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
W-Temper 1/2H-Temper T4-Temper
S
t
r
a
i
n

H
a
r
d
e
n
i
n
g

E
x
p
o
n
e
n
t
n - Compression
n - Rolling
Figure 4-37: Bar graph comparing the strain hardening exponent (n) in cold rolling and
compression relative to the material temper.


155
0
100
200
300
400
500
600
700
800
900
1000
W-Temper 1/2H-Temper T4-Temper
S
t
r
a
i
n

H
a
r
d
e
n
i
n
g

P
a
r
a
m
e
t
e
r

(
M
P
a
)
0
20
40
60
80
100
120
140
S
t
r
a
i
n

H
a
r
d
e
n
i
n
g

P
a
r
a
m
e
t
e
r

(
k
s
i
)
K - Compression
K - Rolling
Figure 4-38: Bar graph comparing the strain hardening parameter (K) in cold rolling and
compression relative to the material temper.


The higher flow stress and SHR in the compression testing relative to the rolling
could be a result of multiple experimental differences between the data sets. These
differences include the mode of deformation, strain rate, homogeneity of the imposed
strain, and frequency of data collection. Regardless of the precise reason why the data for
the different strain paths arent in better numerical agreement, the overall trends in the
two data sets are. It should be noted that it would not be expected for the stress-strain data
for different deformation modes to lie directly on top of each other. Discrepancies in
stress-strain response have been reported by others when comparing different strain paths
in Al alloys [131].

156
The material tested in compression after rolling is subjected to a strain path
change unlike the material tested at zero strain (zero-time). The transient effect of the
change in deformation mode on SHR is of interest because it impacts strain localization.
Strain localization occurs as a result of the decreasing nature the SHR, not the increase in
flow stress during deformation. In every temper the initial SHR increased with increasing
prior rolling strain, as shown in Figure 4-28 through Figure 4-33. This is believed to be a
direct result of the dislocation structure present after rolling and latent hardening. The
newly activated dislocations not only have to interact with the already existing
dislocation forests but also with the stress fields of dislocations on slip systems that are
no longer activated. As the compressive strain increases, the preexisting dislocation
structure is destroyed and the transient effect on SHR ends. The experimental results also
show that solute clustered into GP zones magnify the effect. This could possibly be a
result of dislocations being preferentially activated on certain slip systems as a result of
the coherent nature of the zones. This could create a more ordered dislocation structure
that would subsequently need to be destroyed. Once the effects of the prior dislocation
structure have been eliminated ( 0.07) the material in the W-temper once again most
effectively strain hardens.

4.5.4 Inhibition of GP Zone Formation

It is known that the completion of the natural aging process coincides with the
loss of the excess vacancies produced by quenching [49]. The absence of the excess
vacancies does not alter the driving force for the solute to cluster but rather drastically
alters the kinetics. As shown previously, at room temperature the bulk diffusion rate of

157
solute in an Al matrix is so small that the time frame for GP zones to form would be
orders of magnitude larger than the 1 to 4 days observed. It was also shown in a previous
section that the rate of quenching directly affects the natural aging kinetics, primarily due
to its impact on the resulting concentration of excess vacancies. What is believed to be
inhibiting the continuation of the aging process, seen in Figure 4-16 and Figure 4-17,
after deformation is the accelerated loss of excess vacancies. If a large enough dislocation
density is present in the material, the equilibrium vacancy concentration at room
temperature should be reached more rapidly. This is a result of the dislocations acting as
sinks for vacancies. At a higher dislocation density, the distance that a vacancy has to
travel by random walk to reach a dislocation is decreased, resulting in a higher
probability of annihilation. The long range stress filed of a dislocation may also interact
with the short range stress field of the vacancy producing an attraction between them. It is
known that the stress fields of dislocations and solute atoms interact. The clustering of
vacancies and solute atoms may also result in a coupled motion, increasing the rate of
annihilation. When the mobile vacancy reaches the dislocation it is eliminated by
dislocation climb. The rapid reduction of excess vacancies, in turn, effectively halts GP
zone formation.
In order for the dislocation-vacancy interaction to occur rapidly enough to inhibit
GP zone formation, the dislocation spacing and vacancy spacing must be such that the
vacancies can diffuse to the dislocations in a time frame on the order of seconds to
minutes. The average spacing between vacancies can be estimated using equation 4.6 and
the geometric relationship:

158

3
1
|
|
.
|

\
|
|
|
.
|

\
|
|
|
.
|

\
|
|
|
.
|

\
|
=
Parameter Lattice
Cell Unit
Cell Unit
Site Lattice
Site Lattice
Vacancy C
d
v
v
(4.7)
with 4 lattice sites per fcc unit cell and 4.05E-10m as the lattice parameter for aluminum.
The mean distance a vacancy can travel as a function of time is given by:

( )
2 1
t D L
V V
=
(4.8)
where D
V
is the vacancy diffusion coefficient for three-dimensional random walk
described by the relationship:

v V
zr D
6
1
=
(4.9)
with the jump distance being equal to the lattice parameter and the coordination number
z being equal to 12 for the fcc unit cell. The vacancy jump frequency r
v
is given by:

KT
E
o v
m
e r

=v
(4.10)
where E
m
is the vacancy migration energy, which is 0.65eV [129] for a vacancy in Al,
and
o
is the thermal vibrational frequency which is approximately equal to the Debye
frequency (10
13
sec
-1
). The diffusion distance of a vacancy as a function of time is show
in Figure 4-39 and the average vacancy spacing for each temper is shown in Table 4-4.


159
0
20
40
60
80
100
120
0 100 200 300 400 500 600
Time (sec)
L
e
n
g
t
h

(
n
m
)
Vacancey Diffusion Distance

Figure 4-39: Mean vacancy diffusion length plotted against time at room temperature for
a ten minute interval (equ. 4.8).



Table 4-4: Mean vacancy concentration (equ. 4.6) and spacing at the time of deformation
(equ. 4.7).
Condition C
v

Spacing
(nm)
W-Temper 9.98E-06 12
1/2H-Temper 3.99E-06 16
T4-Temper 2.62E-14 8589


It is clearly seen that vacancies (regardless of the concentration) are extremely mobile
and can travel a distance of over 100nm (~250 lattice sites) within 10 minutes at room
temperature. The vacancies are also in high concentration and spaced less then 20nm
apart during rolling in all conditions except the T4-temper, which is at the equilibrium
concentration and spacing.

160
The average dislocation spacing for dislocations interacting with soft obstacles,
such as shearable precipitates, dislocation stress field interactions in Stage I hardening,
and penetrable dislocation cell walls that start forming during stage II hardening is
obtained through the geometric relationship:

2 1
3
|
|
.
|

\
|
=

S
L
(4.11)
which assumes uniformly dispersed dislocations. The average dislocation spacing for
dislocations interacting with hard obstacles, such as noncoherent precipitates and
dislocation structures containing jogs which start to form during stage II hardening is
given by the relationship:
2 1
1
o
=
H
L
(4.12)
The dislocation spacing can then be related to the shear stress and macroscopic flow
stress through the dislocation density using the following relationships:

t o M =
(4.13)

2 1
o t Gb = A
(4.14)

L
Gb
= t
(4.15)
where M is the Taylor factor, is a material dependent constant, G is the shear modulus,
b is the Burgers vector, and is the dislocation density. The required dislocation density
to produce an equivalent spacing of that of the vacancies at the concentrations predicted
for the W-temper and half hardened material was determined using equations 4.11 and
4.12, as shown in Figure 4-40 and Table 4-5. The variables used to estimate the
dislocation spacing are shown in Table 4-6.

161

0
5
10
15
20
25
30
35
40
45
50
0.0E+00 5.0E+16 1.0E+17 1.5E+17 2.0E+17 2.5E+17 3.0E+17
Dislocation Density (m/m
3
)
S
p
a
c
i
n
g

(
n
m
)
0
500
1000
1500
2000
2500
3000
3500
4000
F
l
o
w

S
t
r
e
s
s

(
M
P
a
)
Dislocation Spacing (Hard)
Dislocation Spacing (Soft)
Vacancy Spacing (W-Temper)
Vacancy Spacing (1/2H-Temper)
Flow Stress

Figure 4-40: Feature spacing and Flow Stress plotted against Dislocation density. Note
that the two Y-axes can only be read against the X-axes and not each other.



Table 4-5: The data plotted in Figure 4-40.
Condition
Vacancy
Spacing (nm)
Flow Stress
(MPa)
Shear Stress
(MPa)
Dislocation Density
(m/m
3
)
W-Temper 12 1060 - 2080 346 - 667 2.85E+14
1/2H-Temper 16 795 -1530 260 - 500 1.55E+14
T4-Temper 8589 < 0.3 < 0.1 5.41E+08



Table 4-6: Variables used to estimate the dislocation spacing.
0.3
b 2.86E-10 m
G 28,000 MPa
M 3.06



162
The flow stresses required to produce comparable vacancy-dislocation spacing are
unrealistic, as the material would reach Stage IV hardening well before these stress levels
were reached. However, that fine of a dislocation spacing is not necessary due to the high
mobility of the vacancies. What is required is for the vacancy to be able to diffuse half
the dislocation spacing in the experimentally observed time frame. In 10min this would
correspond to a dislocation spacing of approximately ~200nm and an increase in flow
stress of ~120 MPa. The stress predicted to rapidly halt natural aging assumes that the
increase in flow stress is from dislocation interaction only. This assumption could
possibly result in the under estimation of the required increase in flow stress necessary to
halt GP zone formation.
The dislocation density predicted for the T4 material is in the range of that present
in a recrystallized (strain free, 1E8 to 1E10 m
-2
) material. An interesting observation is
that when the natural aging process reaches the T4 condition the vacancy spacing and
dislocation spacing is similar. The dislocation spacing in all conditions is orders of
magnitude smaller then the spacing of other dislocation sinks. Although, the other sinks
are 2-D (interfaces) crystal defects while dislocations are 1-D (line).
The sudden drop in the conductivity upon cold rolling is not associated with the
formation of GP zones and is believed to be a result of the newly generated dislocations.
The increase in conductivity following the initial drop is most likely due to solute atoms
segregating to dislocations to relax their stress fields, effectively condensing two electron
scattering points into only one. If this subsequent change in conductivity was a result of
further GP zone formation it would be expected to decrease rather then increase the
conductivity. A slight indication of this behavior is exhibited in the sample rolled to a

163
strain of 0.15 in the W-temper and half hardened condition. After the initial drop in
conductivity, there is a perturbation in the increasing trend where the conductivity drops
as the hardness is increasing. The magnitude of this drop is larger in the material
deformed in the W-temper, as would be expected due to the higher concentration of
vacancies. The conductivity observations support the inhibition of natural aging as a
result of the increased vacancy-dislocation interaction at/or above a strain of 0.2.
The rolling stress-strain data in Figure 4-22 provides the information necessary to
answer the previously posed question of: what amount of strain needs to be imposed to
generate a high enough dislocation density to produce a mean dislocation spacing that
vacancies can diffuse to on a time frame of seconds to minutes. At a true strain of 0.2 the
dislocation spacing is estimated to be between 45nm and 110nm (soft and hard
obstacles), as shown in Figure 4-41. The time it would take for a vacancy to travel half
the distance of the dislocation spacing is estimated to be between 25sec to 35sec. This
time period of less than one minute is in agreement with the experimental results at
strains above 0.2. However, at strains of below 0.2 GP zone formation is predicted to be
halted much faster then the hours to days that were experimentally observed. The
possible explanations for this discrepancy at lower strains are most certainly related to the
assumptions made about the uniformity of the deformed structure. It is being assumed
that both the strain and dislocation structure are uniform throughout the material. This is
known to be untrue, especially at the lower strains imposed by cold rolling. The
dislocation density and spacing could be significantly different from one grain to another
depending on a grains location and crystallographic orientation within the sample. It is
possible that GP zone formation is halted in some more heavily deformed grains and not

164
in others. The vacancy diffusion rate was calculated using variables that were
experimentally determined using high purity aluminum and assume that only mono-
vacancies are present in the structure. It is known that vacancies not only preferentially
interact with dislocations but also with solute atoms and each other. The interaction
between vacancies and solute atoms is responsible for the formation of GP zones at the
vacancy concentrations observed [96]. Vacancy-solute interaction may slow the vacancy
inhalation rate. Vacancies will also combine with each other forming di-vacancies or, at
high enough concentrations for extended time, pores. This would slow the travel rate of a
vacancy.


0
50
100
150
200
250
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40
Rolling Strain
D
i
s
l
o
c
a
t
i
o
n

S
p
a
c
i
n
g

(
n
m
)
W-Temper (Hard)
W-Temper (Soft)
1/2H-Temper (Hard)
1/2H-Temper (Soft)
T4-Temper (Hard)
T4-Temper (Soft)

Figure 4-41: Dislocation spacing for hard (solid lines) and soft (broken lines) obstacles
plotted against rolling strain for the data in Figure 24.



165
What can be concluded from this data is: at strains above 0.2, the dislocation
spacing is small enough and the vacancy diffusion rate is high enough to effectively
accelerate vacancy annihilation, halting GP zone formation. The precise strain required to
inhibit zone formation cannot be determined do to the assumptions made in calculating
the vacancy-dislocation interactions. The possible annihilation of GP zones during the
deformation processing was also not considered. If these same type calculations are
performed for a strain free material the resulting time to stop GP zone formation is on the
order of hours to days which is in reasonable agreement with experimental results.

4.6 Conclusions

Increasing the quench rate results in an increased rate of natural aging, final
hardness, and residual stress contained in the material.
High enough levels of residual stress can be produced during quenching to result
in mid-plane cracking upon cold rolling.
Uphill quenching is an effective method of relieving thermally induced residual
stress and allowed for CR to be performed without mid-plane cracking.
The typical rapid formation of GP zones at room temperature (Natural Aging) is
inhibited by moderate cold rolling strains (
CR
0.15) through dislocation aided
vacancy annihilation.
Strain hardening is more effective in the presence of solute than GP zones. As GP
zones form, the SHR continually decreases.
A change in strain path produces a transient increase in SHR in both the W and
T4 temper rather then the drop observed in other high SFE alloys.

166
4.7 Future Work

The "Effect of GP Zones on Strain Hardening Rate" have been investigated.
Direct measurement and characterization of strength and SHR have been performed.
However, indirect measurements (hardness and electrical conductivity) were used to
observe the natural aging process. A detailed TEM investigation would be the natural
extension of this study. The primary objectives of the TEM study would be to confirm:
(1) the conclusion that 2524 deformed in the W-temper produces a greater dislocation
density than when deformed to an equivalent strain in the T4-temper and (2) that GP zone
formation is inhibited in a deformed microstructure through direct observation. A further
extension of the current work would be to investigate 6xxx and 7xxx series alloys to
determine if they behave in the same manner.
The residual stress imparted during the rapid quenching of the samples presented
an obstacle that may have prevented this study from being completed. The uphill
quenching procedure relieved the residual stress to a level which allowed for the CR to be
performed. The presence of a residual stress in heat treatable aluminum alloys is a
problem the plagues the aluminum applications industry. The uphill quenching process
needs to be further investigated and optimized to determine if it has industrial potential or
if it will remain confined to the laboratory.


167
CHAPTER 5. IMPROVED STRENGTH THROUGH SOLUTE ENHANCED
STRAIN HARDENING


The wrought precipitation hardened aluminum alloys (2xxx, 6xxx, and 7xxx)
exhibit the greatest strengths commercially available for Al based alloys. The strengths
currently achievable in Al alloys are shown in Figure 5-1 [132] and reach a maximum
yield strength value of ~550MPa in the 2xxx series depending on the specific
composition and processing. The SESH portion of the current study has two goals: (1) to
determine if SESH can provide a viable processing route for producing Al alloys with
comparable yield strength as those currently available but with improved tensile
properties (UTS, % elongation, uniform strain, reduction in area) and (2) to determine the
potential for producing material with strengths currently not commercially available.



168
0
100
200
300
400
500
600
700
0 100 200 300 400 500 600 700
Yield Strength (MPa)
U
T
S

(
M
P
a
)
0
10
20
30
40
50
60
70
80
90
100
0 10 20 30 40 50 60 70 80 90 100
Yield Strength (ksi)
U
T
S

(
k
s
i
)
1xxx
2xxx
2x24
3xxx
5xxx
6xxx
7xxx

Figure 5-1: Typical strength of Al alloys [132].



5.1 Solute Enhanced Strain Hardening

Solute Enhanced Strain Hardening is a strengthening mechanism that amplifies
the effect of strengthening by dislocation interactions through inhibiting dynamic
recovery. The significant effect of large amounts of solute in solution on strain hardening
was first discussed by Dorn et al. [3] in 1950, however, it has received little attention in
the field of metallurgy compared to other strengthening methods. A two part paper was
published by Doherty et al. [5, 6] in 1993 which investigated a wide range of Al based
alloys. The first paper [5] reproduced the result reported by others that as the amount of
unprecipitated solute (solute in solution or clustered into GP zones) is increased both the
yield strength and SHR at equivalent strains is increased (Figure 4-1 and Figure

169
4-2).The effectiveness of various alloying elements on strengthening were also ranked
using the empirical relationship:
(5.1)
P
i i Al alloy
X H K K =
where K
alloy
is the strain hardening parameter of the alloy being evaluated, K
Al
is the
strain hardening parameter of pure aluminum, X
i
is the atomic fraction of the alloying
addition, and the solute enhanced strain hardening parameters are H
i
and P, with P being
equal to 2/3. The effectiveness of an alloying element on increasing strain straining
hardening (H
i
) is more clearly show by plotting this relationship on a log-log scale, as
shown in Figure 5-2.

y = 1583x
R
2
= 0.9943
y = 3713.2x
R
2
= 0.9799
y = 4455.3x
R
2
= 0.9763
0
50
100
150
200
250
300
350
400
450
500
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2
X
(2/3)
D
e
l
t
a

K

(
M
P
a
)
0
10
20
30
40
50
60
70
D
e
l
t
a

K

(
k
s
i
)
Al-Cu-Mg (Cu>Mg)
Al-Cu
Al-Mg
Figure 5-2: Delta K (K
alloy
K
Al
) plotted against atomic fraction of solute [5].


170
Examination of Figure 5-2 clearly shows that Cu inhibits dynamic recovery more
effectively then Mg at equivalent atomic percents. The ranking of the elements that were
investigated are shown in Table 5-1.

Table 5-1: Values of solute strengthening parameters for alloying additions [6].
Element H
i
(MPa)
Copper 3,600
Germanium 1,560
Silicon 1,500
Magnesium 1,470
Lithium 1,060
Manganese 1,060
Silver 740
Titanium 630
Gallium 300
Zinc 280


It is also apparent from Figure 5-2 that Cu-Mg (Cu>Mg) solute additions have a
synergetic effect. This synergetic effect is ratio dependent and was not seen in all ternary
alloys. While the first paper concentrated mainly on mechanical properties, the second [6]
examined the microstructures. The local elastic lattice strain was measured by means of
x-ray line breadth for the strain hardened material. A striking feature is that when the
yield strength is plotted against x-ray line breath it revealed a well correlated liner
relationship, as shown in Figure 5-3.


171

{220} X-Ray Line Breadth ()
Y
i
e
l
d

S
t
r
e
n
g
t
h

(
M
P
a
)

Figure 5-3: Yield Strength plotted against X-Ray line Breadth for Aluminum samples
with various amounts of solute for the alloying elements listed in Table 2 [6].



The increase in lattice strain represents an increase in dislocation density which is
the origin of the increased strain hardening. Transmission Electron Microscopy (TEM)
was also used to investigate the dislocation structure of the strain hardened material. It
was reported that the low solute samples contained well defined dislocation cell walls,
separating dislocation free volumes with misorientations of over 15. This result is well
known and has been reported by other investigators. The high solute material that was
deformed exhibited very different dislocation structures. The dislocations did not form
sub-grains. The dislocation structure was uniform throughout the structure with
misorientations of less then 5 within preexisting grains. The non-recovered dislocation
structures observed in the high solute material are in good agreement with the previous
observations of increased yield strength, increased strain hardening, increased lattice
strain, and illustrates the inhibition of dynamic recovery.




172
5.3 Processing and Experimental Method

The experimental procedure for determining if SESH is capable of producing Al
alloys with comparable yield strengths and improved tensile properties to those currently
available is depicted in Figure 5-4. The yield strength-rolling strain information from the
previous experiments, in conjunction with artificial aging curves, will provide the
necessary information to create a series of samples with varying yield strengths.


Solution Heat
Treat
Water Quench
Uphill Quench
W or T4-Temper
Artificially Age Cold Roll
Strain from
Hollomon equ.
UA, T6, OA
Tensile Testing
Microstructure characterization
Data Analyses
Figure 5-4: Flow chart depicting the experimental procedure for the SEHS part of the
current study.






173
5.3.1 Material and Specimen extraction

The AA2524 plate used for the experiments within this section was from the same
stock material as that used for the work in the previous chapter. The starting plate
received from Alcoa in the F-temper was subdivided into eight smaller 114mm x 86mm x
10mm (RD, TD, ND) coupon size plates, after removing the cladding. Half of the coupon
sized plates were artificially aged and the other half were CR in either the W or T4-
temper. Specimens were removed from each of the coupon size plates in the orientations
show in Figure 5-5. It should be noted that the rolling direction in the as-received plate
was held constant throughout the entire study.



Figure 5-5: Schematic representation of a coupon size plate with tensile and C(T)
specimen nested inside so that the loading direction is parallel to the rolling direction.



RD
TD
114 mm
86 mm 86 mm

174
5.3.2 Artificial Aging of AA2524

Four coupon size plates were SHT and uphill quenched in the same manner as
previously described. All four plates were allowed to naturally age at room temperature
to the T4-temper. Upon the completion of natural aging, the specimens were artificially
underaged (UA), peakaged (T6), and overaged (OA) at 190C for the times shown in
Table 5-2. Three reference thermocouples were placed in contact with each sub-size
plate during aging. The time at temperature was started once the last reference
thermocouple reached 190C. The quench delay was less than 5 sec.


Table 5-2: Processing parameters for the aged coupon size plates.
Temper Temp. (C) Time (hr)
T4 ~21 120+
UA 190 4.8
T6 190 10.5
OA 190 21.5


5.3.3 SESH of AA2524

The second four of the eight coupon size plates were SHT and uphill quenched in
the same manner as previously described. Two of the four plates were allowed to
naturally age at room temperature to the T4-temper prior to CR. The other two plates
were CR in the W-temper immediately after the uphill quenching was completed, as
shown in Table 5-3. The largest reductions possible per pass through the rolling mill
were taken to insure uniform strain through the thickness of the plates. Water quenching
between passes through the rolling mill were performed if the material became hot to the
touch.

175
Table 5-3: Processing parameters for the SESH coupon size plates.
Temper CR Strain
T4 0.16
W 0.17
W 0.31
T4 0.41



5.3.4 Microstructure Characterization

Standard metallographic specimen for each of the eight conditions were mounted,
polished, and etched using Keller's reagent. Representative microstructures were selected
in two areas for each of the processing conditions to perform grain morphology analyses.
Grain size and aspect ratios (ND/RD) were measured for each processing condition using
the intercept method. Ten parallel lines were placed in the RD and ND directions in each
micrograph.
Microhardness traverses were taken on each of the eight metallographic
specimen, in accordance with [133], on the RD face in the ND direction. Measurements
were taken every 0.25mm from the centerline to a distance of 3.05mm. A Vickers
indenter was used with a 200g force. Four traverses were taken on each specimen for
each condition.
The through thickness composition of the material was determined using an
Oxford Instruments Energy Dispersive Spectroscopy (EDS) system. Full field of view
compositional measurements were taken such that four microhardness indents where in
view (the outer most two bounding the image). The field of view was then moved three
microhardness indents (one indent overlapped during each scan) toward the centerline
and the next scan was taken. A JOEL JSM-6460LV SEM at working distance of 10mm,
spot size of 40, and accelerating voltage of 20KV was utilized.

176
5.3.5 Tensile Testing

Tensile testing was performed using a MTS servohydrolic load frame in
accordance with [134]. Round sub-size tensile bars, with the dimensions shown in Figure
5-6, were milled from each plate. Extensometer data was collected up to the point of
failure.


Figure 5-6: Schematic showing the dimensions of a the ASTM sub-size tensile bars
extracted from the processed plates.


5.4 Experimental Results

Preliminary experimental results and details that were not directly related to the
conclusions drawn within this document are included in Appendix C and D. These results
include the Equal Channel Angular Pressing (ECAP) of commercial purity aluminum and
the ensuing mechanical properties and crystallographic texture. Appendix C also includes
limited results of the effect of ECAP processing on AA5182 and AA2524. A particularly
notable result reported for AA2524 is that a yield strength of 620MPa was achieved at an
R=4mm
102 mm
4 mm 6 mm
32 mm
32 mm
RD
TD

177
ECAP strain of 1. The remainder of section 5.3 will be devoted to experimental results
that directly support the conclusions in this chapter.

5.4.1 Microstructure Characterization

The grain size measurements for the aged specimen showed a grain morphology
with average sizes of 42m (RD), 40 m (TD), and 18m (ND), as shown in Table 5-3.
The material in the T4 condition exhibited the largest grain size. As the aging time is
increased the average grain size decreases, as shown in Figure 5-7. The grain size trend
is the opposite of what would be expected although within the scatter of the
measurements. The average grain size in the TD direction also appears to be larger than
in the CR direction in the UA specimen, although this is also within the measurement
error.
The grain size measurements for the SESH specimen showed a grain morphology
that varied with the amount of imposed strain. Relative to the average aged values, the
grain size increased in the RD direction and decreased in the ND direction, as shown in
Figure 5-8. This pancaking of the grains is more evident when observing the RD/ND
aspect ratio for each processing condition, as shown in Figure 5-9. Representative
micrographs for each processing condition are contained in Appendix E.





178
Table 5-4: Grain sizes for each processing condition. Data graphically displayed in
Figure 5-7, Figure 5-8, and Figure 5-9.
Grain Size (um)
Temper
RD TD ND
Ratio
RD/ND
T4 59 57 20 3.0
UA 38 42 20 1.9
T6 36 33 16 2.3
OA 33 31 15 2.2
T4, E=0.16 50 42 18 2.7
W, E=0.17 60 36 15 3.9
W, E=0.31 50 32 13 3.9
T4, E=0.41 70 39 12 6.1


42
40
18
0
10
20
30
40
50
60
70
80
T4 UA T6 OA Average (aged)
G
r
a
i
n

S
i
z
e

(
u
m
)
RD
TD
ND

Figure 5-7: Grain size for each of the four aged microstructures and the average of the
four in the RD, TD, and ND directions.

179
42
40
18
0
10
20
30
40
50
60
70
80
90
Average (aged) T4, E=0.16 W, E=0.17 W, E=0.31 T4, E=0.41
G
r
a
i
n

S
i
z
e

(
u
m
)
RD
TD
ND

Figure 5-8: Grain size for each of the four CR microstructures in the RD, TD, and ND
directions and the average of the aged grain sizes.




180
2.3
0
1
2
3
4
5
6
7
T4 UA T6 OA Average
(aged)
T4,
E=0.16
W,
E=0.17
W,
E=0.31
T4,
E=0.41
A
s
p
e
c
t

R
a
t
i
o

(
R
D
/
N
D
)

Figure 5-9: Grain aspect ratio (RD/ND) for each processing condition.



The Vickers microhardness values for the aged plates exhibit "V" shaped profiles
with the surfaces being harder than the centers. The majority of the data points in each
profile are within 5 HV points of the average hardness for the plate, as shown in Figure
5-10. The average hardness increases as aging progresses from the T4 to T6-temper and
then decreases to the lowest average value in the OA condition.



181
115
120
125
130
135
140
145
150
155
160
165
170
175
180
185
190
195
-3.5 -3.0 -2.5 -2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Distance from Centerline (mm)
H
V
1128
1188
1248
1308
1368
1428
1488
1548
1608
1668
1728
1788
1848
1908
S
t
r
e
s
s

(
M
P
a
)
T4
Ave. (140 HV)
115
120
125
130
135
140
145
150
155
160
165
170
175
180
185
190
195
-3.5 -3.0 -2.5 -2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Distance from Centerline (mm)
H
V
1128
1188
1248
1308
1368
1428
1488
1548
1608
1668
1728
1788
1848
1908
S
t
r
e
s
s

(
M
P
a
)
Under Aged
Ave.(146 HV)


115
120
125
130
135
140
145
150
155
160
165
170
175
180
185
190
195
-3.5 -3.0 -2.5 -2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Distance from Centerline (mm)
H
V
1128
1188
1248
1308
1368
1428
1488
1548
1608
1668
1728
1788
1848
1908
S
t
r
e
s
s

(
M
P
a
)
T6
Ave. (152 HV)

115
120
125
130
135
140
145
150
155
160
165
170
175
180
185
190
195
-3.5 -3.0 -2.5 -2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Distance from Centerline (mm)
H
V
1128
1188
1248
1308
1368
1428
1488
1548
1608
1668
1728
1788
1848
1908
S
t
r
e
s
s

(
M
P
a
)
Over Aged
Ave. (132 HV)

Figure 5-10: Vickers microhardness profiles for the aged plates taken on the RD plane in
the ND direction.


The Vickers microhardness values for the SESH plates exhibit a similar "V"
shaped profile with the exception of the plate processed to a true strain of 0.16 in the T4
condition. The majority of the data points in each profile are within 10 HV points of the
average hardness for the plate. Compared to the hardness profiles of the aged plates the
difference between the maximum and minimum HV values per data set have doubled as
has the noise within each data set, as shown in Figure 5-11. The average hardness values
for each traverse is given in Table 5-5.

182
115
120
125
130
135
140
145
150
155
160
165
170
175
180
185
190
195
-3.5 -3 -2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5 3 3.5
Distance from Centerline (mm)
H
V
1128
1188
1248
1308
1368
1428
1488
1548
1608
1668
1728
1788
1848
1908
S
t
r
e
s
s

(
M
P
a
)
T4, E=0.16
Ave. (170 HV)

115
120
125
130
135
140
145
150
155
160
165
170
175
180
185
190
195
-3.5 -3.0 -2.5 -2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Distance from Centerline (mm)
H
V
1128
1188
1248
1308
1368
1428
1488
1548
1608
1668
1728
1788
1848
1908
S
t
r
e
s
s

(
M
P
a
)
W, E=0.17
Ave. (168 HV)


115
120
125
130
135
140
145
150
155
160
165
170
175
180
185
190
195
-3.50 -3.00 -2.50 -2.00 -1.50 -1.00 -0.50 0.00 0.50 1.00 1.50 2.00 2.50 3.00 3.50
Distance from Centerline (mm)
H
V
1128
1188
1248
1308
1368
1428
1488
1548
1608
1668
1728
1788
1848
1908
S
t
r
e
s
s

(
M
P
a
)
W, E=0.31
Ave. (171 HV)
115
120
125
130
135
140
145
150
155
160
165
170
175
180
185
190
195
-3.5 -3.0 -2.5 -2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Distance from Centerline (mm)
H
V
1128
1188
1248
1308
1368
1428
1488
1548
1608
1668
1728
1788
1848
1908
S
t
r
e
s
s

(
M
P
a
)
T4, E=0.41
Ave. (181 HV)

Figure 5-11: Vickers microhardness profiles for the SESH plates taken on the RD plane
in the ND direction.


Table 5-5: Average hardness value and standard deviation for each processing condition.
Temper
CR
Strain
Ave. Hardness
(HV)
STDEV
(HV)
T4 0 140 3
UA 0 145 4
T6 0 152 3
OA 0 132 4
T4 0.16 170 2
W 0.17 168 6
W 0.31 171 8
T4 0.41 180 7


The three main alloying elements in AA2524 are Cu, Mg, and Mn. The EDS
profiles for these elements reveal the that Mg and Mn compositions are constant through

183
the material thickness, as shown in Figure 5-12. The profile for Cu reveals that the
centerline is 5% lower in elemental composition.


0.0
0.5
1.0
1.5
2.0
2.5
3.0
3.5
4.0
4.5
-5.0 -4.0 -3.0 -2.0 -1.0 0.0 1.0 2.0 3.0 4.0 5.0
Distance from Centerline (mm)
W
e
i
g
h
t

%
Copper (Cu)
Magnesium (Mg)
Manganese (Mn)

Figure 5-12: EDS through thickness elimental composition profile for the material in the
T4-tmeper.


5.4.2 Tensile Data

The strength data for the aged material increases from the T4-temper to the UA
condition until reaching the T6-temper. Further aging from the T6-temper to the overaged
condition results in a decreases in strength, as shown in Figure 5-13. The strength of the
SESH material increases with increasing strain. The greatest strength exhibited for the
CR material was for the plate processed to a true strain of 0.41 in the T4-temper and the

184
lowest strength was observed in the plate processed to a true strain of 0.17 in the W-
temper, as shown in Table 5-6. Typical engineering stress-strain curves for each
processing condition are shown in Figure 5-14.


300
350
400
450
500
550
600
300 350 400 450 500 550 600
Yield Strength (MPa)
U
T
S

(
M
P
a
)
43.5
48.5
53.5
58.5
63.5
68.5
73.5
78.5
83.5
43.5 48.5 53.5 58.5 63.5 68.5 73.5 78.5 83.5
Yield Strength (Ksi)
U
T
S

(
K
s
i
)
T6 to OA
T4 to T6
CR - W
CR - T4
Linear (CR Data)

Figure 5-13: Ultimate tensile strength (UTS) plotted against yield strength for all eight
processing conditions.

185
0
100
200
300
400
500
600
0.00 0.05 0.10 0.15 0.20 0.25 0.30
Eng. Strain
E
n
g
.

S
t
r
e
s
s

(
M
P
a
)
0
10
20
30
40
50
60
70
80
E
n
g
.

S
t
r
e
s
s

(
K
s
i
)
T4-Temper
UA
T6-Temper
OA
T4, E=0.16
W, E=0.17
W, E=0.31
T4, E=0.41

Figure 5-14: Engineering stress plotted against engineering strain.



Table 5-6: Average tensile properties for all eight processing conditions.
Yield
Strength
(MPa)
UTS
(MPa)
Elongation
(%)
Reduction
in Area (%)
Area
Under
Curve
(J/m^3)
Stain
Hardening
Exponent - n
Strain
Hardening
Parameter - k
(MPa)
T4 0 313 481 20% 42% 122.4 0.207 799
UA 0 360 494 17% 40% 107.0 0.172 781
T6 0 400 493 11% 33% 70.8 0.132 740
OA 0 333 423 13% 44% 65.5 0.150 673
T4 0.16 463 537 12% 38% 82.9 0.098 741
W 0.17 442 519 12% 40% 81.6 0.108 730
W 0.31 465 540 12% 36% 88.7 0.101 747
T4 0.41 500 566 9% 39% 62.6 0.091 786
T
e
m
p
e
r
C
R

S
t
r
a
i
n
Average Values



The inverse relationship between strength and percent elongation is observed for
both the aged and SESH material. The reduced strength in the OA condition does not
result in a recovery in ductility that would be expected at equivalent strengths before peak

186
aging. The CR material exhibits higher yield strengths at equivalent percent elongations,
as shown in Figure 5-15.


0%
2%
4%
6%
8%
10%
12%
14%
16%
18%
20%
22%
24%
300 350 400 450 500 550 600
Yield Strength (MPa)
E
l
o
n
g
a
t
i
o
n

(
%
)
43.5 48.5 53.5 58.5 63.5 68.5 73.5 78.5 83.5
Yield Strength (Ksi)
T6 to OA
T4 to T6
CR - W
CR - T4
Linear (CR Dara)

Figure 5-15: Percent elongation plotted against yield strength for all eight processing
conditions.


The reduction in area (R.A.) data had the greatest amount of scatter. The aged
material exhibited the expected decrease in %R.A. with increasing yield strength, as
shown in Figure 5-16. Unlike the elongation data, when the yield strength decreases in
the OA condition the R.A. recovers to a value equivalent or greater than what would be
expected for the below peak aged strength value. The SESH material does not exhibit a
clear trend, although the strengths appear to be greater at equivalent R.A. values.

187
0%
5%
10%
15%
20%
25%
30%
35%
40%
45%
50%
300 350 400 450 500 550 600
Yield Strength (MPa)
R
e
d
u
c
t
i
o
n

i
n

A
r
e
a

(
%
)
43.5 48.5 53.5 58.5 63.5 68.5 73.5 78.5 83.5
Yield Strength (Ksi)
T6 to OA
T4 to T6
CR - W
CR - T4
Linear (CR Data)

Figure 5-16: Reduction in area plotted against yield strength for all eight processing
conditions.


The area under the engineering stress-strain curve decreases as the yield strength
increases from the T4 to T6-temper. As the yield strength decreases in the OA condition,
there is no additional energy absorbed by the tensile bar, as shown in Figure 5-17. The
SESH material also exhibits a decrease in area under the curve with increasing yield
strength, although the yield strength values are greater at equivalent levels of energy
absorption.

188
0
20
40
60
80
100
120
140
300 350 400 450 500 550 600
Yield Strength (MPa)
A
r
e
a

U
n
d
e
r

C
u
r
v
e

(
J
/
m
^
3
)
0
5
10
15
20
25
30
35
43.5 48.5 53.5 58.5 63.5 68.5 73.5 78.5 83.5
Yield Strength (Ksi)
A
r
e
a

U
n
d
e
r

C
u
r
v
e

(
i
n
-
l
b
f
/
f
t
^
3
)
T6 to OA
T4 to T6
CR - W
CR - T4
Linear (CR Data)

Figure 5-17: Area under the engineering stress-strain curve plotted against yield strength.


5.5 Discussion

5.5.1 Microstructure Characterization

The grain size measurements had a larger variation than anticipated given that all
the coupon size plates were extracted from the same starting stock material. It was
expected that all the aged material would have equivalent average grain sizes. The
thermal history of the artificially aged plates were all similar with the exception of the
aging times. The pinned out grain size produced by the SHT process (520C-525C)
would not be expected to grow any further during the lower temperature (190C) aging
process. Growth of the particle limited grain size produced at the SHT temperature would
be further inhibited by additional precipitation during aging. If the aging time were to

189
have an effect on the grain size (which it does not) the effect that would be expected is
that the grain size would increase as the time at temperature increased. However, the
average measured grain sizes follow the inverse relationship, although the average values
are within the error of the measurements. The material in the T4 condition had the largest
grain size although it was not exposed to the 190C artificial aging treatments. The
increased grain size may be a result of the T4 plate being exposed to a second SHT. The
T4 material was re-solutionized because it spent 5min in the boiling water during the first
uphill quench rather than the intended 1min exposure. The 520C to 525C temperature
of the SHT process is high enough to result in an additional increment of grain growth.
The grain morphology of all the aged microstructures is typical of a rolled and
recrystallized structure in Al with the largest grain direction being the RD and the
smallest being the ND. The microstructure of the UA material exhibited a larger grain
size in the TD direction than the RD. The difference in TD and RD grain sizes in the UA
material are within the noise of the measurements. A reason for the RD and TD grain
sizes in the aged material being within ~10% of each other may be that the stock plate
was cross rolled at Alcoa during the processing. Any scatter in the grain sizes of the aged
material is believed to be representative of the starting stock and a result of industrial
processing performed prior to receiving the material.
The average aged grain size is a valid approximation for the starting grain size
before the CR process. The CR process can be approximated as plane strain deformation
and the evolution of the grain sizes in the SESH material reflect this. As the level of
imposed strain increases the RD grain size increases, the ND grain size decreases and the
both the RD/ND and RD/TD ratios increase. The difference in the Hall-Petch effect

190
between the W-temper and T4-temper has not been investigated and could potentially
play a roll in the relative strengthening of the two conditions.
The average HV value of the OA material is the lowest of the aged plates,
although it does not have the lowest strength. The reason for this is that hardness is a
function of both yield strength and SHR. While the yield strength of the OA material is
higher than that of the T4 material, the SHR is not as high. The strain hardening exponent
of the T4 material is ~40% higher than the OA material which offsets the higher yield
strength.
The microhardness traverses were taken to asses the uniformity of the imposed
strain. The microhardness profiles of the aged material were intended to be used as the
baseline data to compare the SESH material to. The "V" shape of the profiles is not
believed to be a by-produced of the testing but rather representative of the material. The
grain size of the material is uniform through the thickness of the plates therefore the "V"
shape is not expected to be a result of the Hall-Petch effect or texturing of the material.
The "V" shape is also most likely not due to unrelaxed residual stress from quenching as
it would be expected that the state of the residual stress would invert the "V" shape.
The origin of the "V" shape is most likely due to the solute gradient through the
thickness of the plate that formed during the ingot casting process [135, 136].
Microsegregation during casting is not an uncommon occurrence. The first liquid to
solidify is of higher purity creating a solute enriched melt. As the solidification process
proceeds the solute concentration at the solidification interface reaches a steady state and
the magnitude of the solute gradient in the solidified material diminishes.
Homogenization of the material aids in reducing the local composition gradients but not

191
those requiring long range diffusion. The EDS data showed, Figure 5-12, that the Cu
level in the center of the plate was ~5% lower than in the bulk. The Cu is the only
alloying element that the microsegregation is seen in due to the resolution of the EDS and
level of alloying. It would be expected that the microsegregation of Cu would the greatest
due to it having a lower partitioning coefficient (k) than Mg or Mn, as shown in Table
5-7. Where the partitioning coefficient is the ratio of the slopes of liquidus line to the
solidus line on a binary phase diagram. The lower the partitioning coefficient the more
likely an element is to exhibit segregation. The lower Cu level would result in less
effective solution hardening, precipitation hardening, and SESH. The less effective
hardening in the mid-plane would result in lower HV values.

Table 5-7: Partitioning coefficients (k) for various aluminum binary alloy systems [137].
Alloy System k
Al-Fe 0.02
Al-Si 0.14
Al-Cu 0.17
Al-Mg 0.43
Al-Sc 0.52
Al-Mn 0.95


The increase in the magnitude to the "V" shaped HV profile in the SESH material
is not believed to be due to non-uniform strain. There was not a grain size gradient
observed in the SESH material. The increase in the magnitude to the "V" shaped HV
profile is believed to be a result of the CR process increasing the solute gradient and the
decreasing effectiveness of the SESH with decreasing Cu content. The increased noise in
the SESH HV values may be a result of non-uniform strain within grains (not to be
confused with non-uniform strain through the cross-section).

192
5.5.2 Tensile Results

The aging times selected were intended to produce an underaged and overaged
condition of equivalent yield strengths in addition to a peak aged condition. Artificial
aging curves for AA2024 were used to select the aging temperature of 190C since it
provided the greatest peak strength [117]. The aging behavior of AA2524 was expected
to be similar to that of AA2024. However, AA2524 didn't reach the strengths predicted
for AA2024 for the same heat treating procedures, as shown in Figure 5-18. The
measured yield strengths in the AA2524 are 10-20% lower.


0
100
200
300
400
500
600
1 10 100 1000
Time at 190C (h)
S
t
r
e
s
s

(
M
P
a
)
0
10
20
30
40
50
60
70
80
S
t
r
e
s
s

(
K
s
i
)
Yield strength (AA2024)
UTS (AA2024)
UTS-YS (AA2024)
Yield Strength (AA2524)
UTS (AA2524)
UTS-YS (AA2524)
Target Values

Figure 5-18: Stress plotted against aging time at 190C for AA2024 and AA2524. Solid
lines are the aging response of AA2024. Broken lines are trend lines through the
experimentally measured AA2524 values.




193
The lower yield strength values were not a failure of the heat treatment, but rather
a less effective response of the alloy to the heat treatment than expected. The strength for
the aged material fell within the expected range [132] for 2x24 alloys, as shown in
Figure 5-19. The AA2524 did progress from the underaged condition to peak aged and
then to the overaged condition. The less effective precipitation strengthening of the
AA2524 may be a result of the tighter composition ranges of Cu, Mg, and, Mn.

0
100
200
300
400
500
600
0 100 200 300 400 500 600
Yield Strength (MPa)
U
T
S

(
M
P
a
)
0
10
20
30
40
50
60
70
80
0 10 20 30 40 50 60 70 80
Yield Strength (ksi)
U
T
S

(
k
s
i
)
2xxx
2x24
2524 - Aged
2524 - SESH

Figure 5-19: Ultimate tensile strength plotted against yield strength for 2xxx series alloys.






194
The mechanical properties of the aged material are the baseline to which the
SESH material is being evaluated against. The intent of the SESH processing was to
produce material with:
equivalent yield strengths as the precipitation hardened material (target
values of 400MPa and 460MPa);
a yield strength above what is currently achievable by conventional heat
treating procedures for AA2x24 (target of 500MPa);
equivalent yield strengths after being deformed in the W and T4
conditions (target of 460MPa).
The Hollomon relationship for the CR material was used to estimate the processing strain
necessary to produce the desired yield strengths. The continuous compressive flow stress
of the material and the compressive yield strength-CR strain relationship have been
determined, as shown in Figure 5-20. It should be noted that the discrepancy between the
compressive flow curves and the CR curves is a result of the strain path change that was
discussed in previous sections. The strain that could be imposed on the SESH material
was limited by practical constraints. The minimum amount of CR strain imposed was
restricted to 0.1 to ensure that the full thickness of the material was being deformed. The
maximum amount of CR strain was limited to ~0.4 to allow for the extraction of test
specimen that would meet all ASTM specifications.

195
y = 569.55x
0.1941
R
2
= 0.9988
y = 515.38x
0.07
R
2
= 0.9761
0
100
200
300
400
500
600
700
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45
True Strain
T
r
u
e

S
t
r
e
s
s

(
M
P
a
)
0
20
40
60
80
100
T
r
u
e

S
t
r
e
s
s

(
K
s
i
)
Compression Curve - W
Compression Curve - T4
CR Compression - W
CR Compression - T4
CR Tensile - W
CR Tensile - T4
Power (CR Compression - W)
Power (CR Compression - T4)

Figure 5-20: True stress plotted against true strain. Solid lines represent continuous
compression testing data. Compressive yield strength-CR strain is represented by discrete
solid data points with broken trend lines. The equations for the trend lines are given.
Tensile yield strength-CR strain is represented by hollow data points and has not been
fitted to trend lines.


The strain imposed during CR is not continuous, but rather discrete increments for
each pass through the rolling mill. The discrete nature of the process combined with the
need to take large enough reductions to deform the entire cross section during each pass
made achieving an exact strain level difficult. However, the target yield strength values
were achieved within 10MPa with the exception of the lowest strength value which was
over strengthened by 42MPa.
The SESH of AA2524 is so effective that it was impractical to produce yield
strengths that were equivalent to those of the aged material. The SESH material with the
lowest yield strength is 11% stronger than the peakaged material and with a 5% greater

196
ultimate tensile strength, as shown in Table 5-8. The increase in UTS with increasing
yield strength for the SESH material is a result of the SHR of the materials. Although the
SHR in the SESH is lower (after being CR) than in the aged material, it diminishes less
rapidly, as shown by the slopes in Figure 5-21. The more rapid decline of the strain
hardening exponent in the aged material is a result of more extensive dynamic recovery.
Dynamic recovery is inhibited in the SESH material.

Table 5-8: Tensile properties for the peakaged material (T6-temper) and the SESH
material processed in the W-temper to a CR strain of 0.17.
Material Condition
Tensile Property
T6 W, E=0.17
Percent
Increase
Yield Strength
(MPa)
400 442 11%
Tensile Strength
(MPa)
493 519 5%
Elongation
(%)
11% 12% 1%
Reduction in Area
(%)
33% 40% 7%
Area Under Curve
(J/m^3)
70.8 81.6 15%


197
y = -0.0009x + 0.4778
y = -0.0003x + 0.238
y = -0.0003x + 0.2279
0.00
0.05
0.10
0.15
0.20
0.25
300 350 400 450 500 550 600
Yield Strength (MPa)
S
t
a
i
n

H
a
r
d
e
n
i
n
g

E
x
p
o
n
e
n
t

-

n
43.5 48.5 53.5 58.5 63.5 68.5 73.5 78.5 83.5
Yield Strength (Ksi)
T6 to OA
T4 to T6
CR - W
CR - T4
Linear (CR Data)

Figure 5-21: Strain hardening exponent (n) plotted against tensile yield strength for
AA2524 in all eight processing conditions.



The percent elongation of the SESH material is equivalent to that of the peak aged
material but with a 11% higher yield strength. The increased ductility of the SESH
material is more clearly shown in the reduction of area data. All of the SESH data sets
exhibit higher R.A. values than the peak aged material with yield strengths ranging from
11% to 25% higher. R.A. is not only a measure of ductility, but also an indicator of
toughness. When strain localization occurs, and a round tensile bar necks, the state of the
stress in the necked region is similar to that at the tip of a crack (triaxial). The similarity
in the stress states at the time of fracture leads to a positive correlation between R.A. and
toughness. Another indicator of toughness is the area under the engineering stress-strain
curve. The area under the curve is a measure of the energy per unit volume absorbed by

198
the material during the test. All but the highest strength SESH material absorbed more
energy before fracture than any of the aged material and at higher strength levels. The
origin of the improved tensile properties at equivalent or higher yield strengths is a result
of the increased SHR at equivalent flow stresses.

5.6 Conclusions

The SESH process is sensitive to composition gradients within the material being
processed.
SESH is capable of producing strengths equivalent and greater to those achieved
through precipitation hardening.
Aluminum alloy 2524 strengthened through SESH to a yield strength 11% greater
than that in the T6-temper exhibits: 5% greater UTS, 1% greater elongation, 7%
greater R.A., and absorbs 15% more energy during tensile testing.
SESH appears to be a viable processing route for producing AA2524 with
improved tensile properties.


5.7 Future Work

In the current study it was not possible to reliably produce overlapping yield
strengths for the SESH and aged material. The reasons that yield strengths in the SESH
material could not be produced in the range of that of the T4 to T6-temper are that: (1)
the alloy being studied so effectively SESHs and (2) the deformation process used to

199
impose the strain combined with the cross sectional thickness of the plates resulted in too
large of discrete strain increments.
An extension of the current work would be to select a deformation process that
allowed for small increments of strain to be more reliably imposed on the material. A
viable method for imposing smaller increments of strain may be through the use of a
plane strain compression fixture. This would allow for through thickness strains to be
imposed at varying increments while simulating the CR process.
The corrosion properties of the SESH are well worth investigating. The corrosion
susceptibility of high strength aluminum alloys is partially a result of the galvanic
coupling between the matrix and precipitates. An example of this is the higher corrosive
susceptibility on the short-traverse plane of various product forms due to the larger area
fraction of precipitates as a result of the increased grain boundary length. The galvanic
coupling of the matrix and precipitates is absent in the SESH material. The absence (or
lower volume fraction) of the precipitates may result in an improved corrosion resistance.



200
CHAPTER 6. EFFECT OF SOLUTE ENHANCED STRAIN HARDENING ON
TOUGHNESS


The wrought precipitation hardened aluminum alloys (2xxx, 6xxx, and 7xxx) are
most commonly utilized in applications where high specific strength-toughness
combinations are required. The strength-toughness combinations currently commercially
available for these alloys are shown in Figure 6-1 [132]. The strength levels for the 2xxx
alloys range from 330MPa to 550MPa and the toughness values range inversely from
44MPa m
1/2
to 14MPa m
1/2
. The goal of the current work is to produce alloys that: (1)
fracture by transgranular dimpled rupture rather than grain boundary ductile fracture and
(2) to achieve combination of strength-toughness that are currently unobtainable.




201
0
5
10
15
20
25
30
35
40
45
50
300 350 400 450 500 550 600
Yield Strength (MPa)
K
I
C

(
M
P
a

m
1
/
2
)
0
5
10
15
20
25
30
35
40
45
43.5 48.5 53.5 58.5 63.5 68.5 73.5 78.5 83.5
Yield Strength (ksi)
K
I
C

(
k
s
i

i
n
1
/
2
)
2xxx
7xxx
Linear (2xxx)
Linear (7xxx)

Figure 6-1: Yield strength plotted against fracture toughness for wrought precipitation
harden aluminum alloys [132].


6.1 Effect of Grain Boundary Precipitates on Toughness
As discussed in the Background (section 2.6) of this document, it is known that
alloy cleanliness, grain morphology, particle spacing, and processing history effect the
fracture toughness of an alloy. The specific role of grain boundary precipitates on
toughness was reviewed and investigated by [76]. The investigation focused primarily on
an Al-Li alloy (Al-10.7%Li-0.22%Mn). The authors used a series of heat treating
procedures to produce a range of yield strengths with varying area fractions of grain
boundary precipices. The effect of the processing on toughness is shown in Figure 6-2.
It should be noted that plane-strain conditions were not achieved in the toughness testing,
although the sample geometry was held constant.

202
0
20
40
60
80
100
120
140
50 100 150 200 250 300
Yield Stress (MPa)
S
l
o
w

B
e
n
d

C
h
a
r
p
y

T
o
u
g
h
n
e
s
s

(
M
P
a

m
1
/
2
)
0
20
40
60
80
100
120
7.3 12.3 17.3 22.3 27.3 32.3 37.3 42.3
Yield Stress (Ksi)
S
l
o
w

B
e
n
d

C
h
a
r
p
y

T
o
u
g
h
n
e
s
s

(
K
s
i

i
n
1
/
2
)
AQ to PA
PA + Reverted + aged
OA + Reverted + aged
PA to OA

Figure 6-2: Variation in slow bend Charpy toughness with yield strength for various heat
treatments [76].


The typical increase in strength and decrease in toughness was observed as the
material aged from the as-quenched (AQ) condition to the peak aged (PA) condition.
During the aging process (green squares), both the strengthening ' precipitates and GB
precipitates were reported to have coarsened. The material in the PA condition was
reverted to dissolve the strengthening ' precipitates while leaving the GB precipitates
relatively unaffected. The strength of the material is reduced back to a level comparable
to the AQ condition; however, the toughness is not recovered. The magnitude of the
unrecovered toughness is believed to be the effect of the still present GB precipitates.
Upon aging again (orange diamonds), the strengthening ' phase re-precipitates and the
preexisting GB precipitates further coarsen. A similar loss of toughness with increasing

203
strength trend is observed, but at a lower level due to the higher area fraction of GB .
The reverted PA material (blue circles) behaved in the same manner.
In order to more directly observe the effect of the GB precipitates on toughness a
second serious of specimen were produced. The investigators aged the Al-Li alloy at a
fixed temperature for a series of times to produce material with a range of GB precipitate
area fractions. All the specimen were then reverted to dissolve the strengthening
precipitates and re-aged together at the same time and temperature. This procedure
produced a series of specimen with equivalent yield strengths (between 158MPa and
171MPa) but with varying area fractions of GB precipitates. The variation in GB
precipitates had a substantial effect on the measured toughness, as shown in Figure 6-3.
What is evident from this study is that there is a great potential to improve the strength-
toughness combination in precipitation hardened aluminum alloys if GB precipitates can
be eliminated.


Figure 6-3: Variation of Charpy toughness with areal fraction of grain boundary , at
constant matrix yield strength [76].



204
6.2 Processing and Experimental Methods

The material used to evaluate the fracture toughness of the SESH AA2524 was
from the same stock used to conduct the experiments in Chapters 4 and 5 of this
document. The processing of the material was performed concurrently with that described
in section 5.3.

6.2.1 Material and Specimen Extraction

Three compact tension specimen were extracted from each of the eight coupon
size plates, as shown in Figure 5-5. The C(T) specimen were extracted from each plate
post-processing. Therefore, the processing for the toughness specimen is identical to that
detailed in section 5.3. The loading direction of the compression specimen, tensile bars,
and the C(T) specimen were all parallel to the RD. The C(T) specimen were extracted
such that they were tested in the L-T (loading direction-crack growth direction)
orientation. The extraction was done utilizing a wire EDM and the C(T) faces were
precision ground. The nominal specimen dimensions are given in Figure 6-4 and Table
6-1. The actual as-machined/tested dimensions were used for all calculations.


205

Figure 6-4: Schematic of a compact tension specimen with nominal dimensions (units of
millimeters). Specimen is in the L-T orientation.


Table 6-1: Nominal dimensions of the compact tension specimen.
Specimen Dimension Length (mm)
Width W 25.4
Height H 30.48 (1.2W)
Thickness B 6.35 (0.25W)
Net Thickness B
N
5.08 (0.8B)
Crack-Starter Notch a
EDM
6.858 (0.27W)
Fatigue Notch a
o
15.49 (0.61W)
Loading Hole D 4.775 (0.188W)


6.2.2 Fracture Toughness Testing

The toughness of the eight processing conditions was characterized by
constructing J-integral resistance curves (J-R curves). The J-R curves were used to
determine the critical J
IC
values, which were subsequently converted to the critical stress
intensity values, K
JIC
. Toughness testing was performed in accordance with [138]
Thickness = 6.35mm
Center line
Load line
31.75 0.254
2.54
5.08
9.068
6.35
30
D=4.775
0.254
2.54
30.48 0.254
RD
TD

206
utilizing an MTS servohydraulic test frame, MTS TestStar II controller, and crack
opening displacement (COD) gauge in conjunction with a Fracture Technology
Associates (FTA) Non-Linear Fracture Toughness interface and digital signal processing
system. The steps below detail the specifics of the toughness testing.
a) Fatigue Pre-cracking: Fatigue pre-cracking was performed using a MTS
servohydraulic test frame, MTS controller, and COD gauge in conjunction with a
FTA Fatigue Crack Growth interface and digital signal processing system.
Traveling microscopes were used to monitor the crack length and stainless steel
shim stock was periodically inserted between the clevis grips and loading pins to
ensure a straight crack front. Pre-cracking was conducted in laboratory air at 8 to
12Hz and an R (K
min
/K
max
) value of 0.1. A load of 2046 N was used to initiate the
crack. After crack initiation the max load was gradually reduced from 1690 to
1070N. The final ~1.5mm was fatigued at 800 to 660N. The ASTM test method
specifies that the K
max
during the final stages of pre-cracking shall not exceed
60% of K
IC
. At no point after crack initiation did K
max
exceed 40% K
IC
. The
nominal fatigue pre-crack length was 8.6mm.
b) Side-Grooving: Once fatigue pre-cracking was completed the specimens were
side-grooved using a wire EDM, as shown in Figure 6-5. The depth of each side-
groove was 10% of the total specimen thickness resulting in a net thickness of
80%. The angle of the side-grooves were 60, as shown in Figure 6-6. Side-
grooving is done after pre-cracking to aid in producing a straight crack front.



207

Load Line
Graphite
Side Groove
TD
RD
Figure 6-5: Side-grooved compact tension specimen with dry graphite lubricant.




Figure 6-6: Compact tension specimen showing net thickness and side-groove angle.

Net Thickness - B
N

60
ND
RD

208

c) Specimen Seating: The specimen was loaded back into a servohydraulic test
frame fitted with D-grip clevises (rather than round grip clevises used for pre-
cracking) using 3.912mm (#23 drill bit blanks) loading pins. Care was taken to
ensure that the pins could roll freely on the loading flats. Graphite was applied to
the bearing surfaces inside the specimen loading holes and on the loading flats of
the clevis. The specimen must fit within the clevis as closely as practically
possibly without contacting the clevises to ensure that there is no bending in the
loading pins. Extreme care was taken to ensure that the loading line of the
clevises were aligned both axially and centrifugally. The alignment of the test
frame/tooling is critical. Before engaging the test system it was verified that all
sensors (load cell, COD, LVDT) were calibrated and within the appropriate
ranges.
d) System Check: A serious of at least three elastic loadings and unloadings were
performed to ensure that the specimen and clip gauge are properly seated, verify
that the sensors were reading correctly, and verify that the initial crack length was
as expected (relative to the crack length calculated using the compliance
relationship). In order to continue on to the actual development of the J-R curve
the three predicted crack lengths must not differ by more than 0.003W, Ci/Co
shall not differ by more than 0.003, and the correlation coefficient for the loading
and unloading slopes must be 0.9999 or better. The purpose of these criteria are to
ensure that the crack lengths/extensions measured during the testing are accurate.
If any of these criteria were not met the procedure was started over at the previous
step. The loading-unloading slopes were taken for each specimen between 355

209
and 890N. Once the above criteria were met, the specimen was not fully unloaded
again until the test was complete.
e) Specimen Testing: The compliance method was used create the J-R curves. The
test procedure entails loading the specimen in displacement control (LVDT
displacement not COD displacement). The loading was systematically interrupted
and partially unloaded to updated the crack length using the specimen
compliance, as shown in Figure 6-7. The data acquisition system continuously
collected load, displacement, and COD values, which allowed for a J value to be
calculated each time the crack length was updated. The test parameters are
provided in Table 6-2. The specimens were loaded at a rate of 0.127mm/min until
either the COD displaced 0.013mm or there was an increase in load of 110N.
Every additional 0.013mm or 110N an unloading slope was determined. The
specimen was unloaded 535N at a rate of 0.254mm/min and held for 1 second.
The slope was calculated using the lower 85% of the unloading. After a duration
of 1sec the specimen was reloaded. The loading-unloading process was continued
until the COD had displaced a total of 2.032mm. Once a COD displacement of
2.032mm is reached the specimen was unloaded at a rate of 0.508mm/min and the
test was complete.


210

Figure 6-7: Load plotted against COD displacement for the unloading used to update the
crack length (T6 specimen). Note that the loading data is not plotted.


211

Table 6-2: Non-linear fracture toughness testing parameters. Specimen specific fields
have been left blank.
Test Machine: C(T)
Temp (C): L-T
% RH: Polynomial
K1 4.64 C1 -4.0632
K2 -13.32 C2 11.242
K3 14.72 C3 -106.04
K4 -5.6 C4 464.33
K5 0 C5 -650.68
Sensor Unit/Volt Full Range Sensor Lower Upper
Load Cell (lb/v) 100 1,000 lbf Load (lbf) -100 975
Disp1 (in/v) 0.005 0.050 in Disp1 (in) -0.01 0.05
Disp2 (in/v) 0.005 Not Active Disp2 (in) -- Not Active
Aux (units/v) 1 Not Active Aux (units) -- Not Active
Stroke (in/v) 0.025 0.250 in Stroke (in) 0.025 0.250 in
0.200
25
0.0005 200
-- --
-- 0.08
120 --
85 --
1
Ramp Rate Acq. Rate
0.005 4
0.01 0.2
0.02 10
--
Spec. Width (in) Tensile Stress (ksi)
Span to Width Ratio
Crack Length (in) Specimen dimensions and material
properties are specimen specific. All other
test parameters are held constant. Fields
populated with " -- " are unused.
Pin Spacing (in)
Comp Alpha Ratio
Spec. Thickness (in) Mod. Of Elasticity (Msi)
Net Spec.Thickness (in) Yield Stress (ksi)
Ramp Down (in/min) Unload (pts/sec)
Specimen Dimensions Material Properties
Ramp Up (in/min) Normal (pts/sec)
Unload (in/min) Hold (pts/sec)
Unload slope delay (sec)
Ramp Generator Data Acq Rate
Command Command
Load drop for slopes (lb) Ramp Down Amount (lb)
Upper window (%) Automatic Preload (lb)
Disp2 Interval (in) Load Drop Limit (lb)
Time Interval (sec) Delta a Limit (in)
Load Interval (lb) Program Limits and Actions
Disp1 Interval (in) Load Limit Hold (lb)
Calibration Sensor Limits
Comp Gage Length (in)
Unloading Slope
MT Lab %RH K Calibration Type:
Stress Intensity Coefficients LL Compliance Coefficients
NLFT Test Parameters
HVHT-2 Specimen Geometry:
MT Lab Temp. Orientation:


212
f) Fatigue Marking: The end of the specimen tearing must be marked by fatiguing
(fatiguing was selected rather than heat tinting so that fractography could be more
effectively performed on the fractured surfaces). After removing the specimen
from the test frame being used to perform the NLFT testing, the specimen was
taken back to the test frame used to perform the pre-cracking. It is important not
to alter the configuration of the test frame being used to perform the NLFT testing
until all testing is completed. The specimen were fatigue marked using the typical
round hole clevises and 4.394mm loading pins (#17 drill bit blanks) in load
control. The max load used was 600N with R=0.1. The specimens were allowed to
run-out until the critical crack length was reached.



a
P

Knife edge, on LL
a
EDM
a
O

Fatigue Pre-crack Tearing Overload
TD
ND
Figure 6-8: Half of a fractured C(T) specimen showing (left to right) the knife edge on
the load line that the COD was mounted on, EDM starter notch, fatigue pre-crack, ductile
tearing from testing, fatigue marker region, and the mechanical overloading when the
critical crack length was reached.



213
g) Physical Crack Length Determination: After the specimen was fractured into two
separate pieces, it was possible to accurately measure the initial (a
o
) and final (a
P
)
crack lengths. Crack length measurements were taken using an optical
comparator. A serious of nine measurements were taken across both the fatigue
pre-crack front and at the end of the ductile tearing. Each series of measurements
were evenly spaced starting and ending at 0.005W in from the root radius of the
side-grooves, as shown in Figure 6-9.

0
4
8
12
16
20
24
28
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5
Net Thinkness - B
N
(mm)
W
i
d
t
h

-

W

(
m
m
)
0.0
0.2
0.4
0.6
0.8
1.0
0 0.05 0.1 0.15 0.2
Net Thinkness - B
N
(in)
W
i
d
t
h

-

W

(
i
n
)
Load Line EDM Starter Notch
Fatigue Pre-Crack End of Ductile Tearing
Specimen Boundary

Figure 6-9: C(T) specimen width plotted against the net thickness showing a serious of
nine measurements for the load line, EDN starter crack, fatigue pre-crack front, and the
end of the ductile tearing (beginning of the fatigue marker). Note that the first and last
measurements are 0.005W in from the root radius of the side-grooves.



214
The initial (a
o
) and final (a
P
) crack lengths were determined by averaging the two
measurements closest to the side-groove root radius (the 1
st
and 9
th
value) to be
used as a single value and then averaged with the remaining seven values, as
shown in equation 6.1 for the initial crack length:

( )
8
2
8
2
9 1
i
o
a
a a
a
E +
|
.
|

\
| +
= (6.1)
If any of the individual crack length measurements differ form the average value
by more than 5% the test data for the specimen is invalid.
h) Calculation of J values: The J value needs were calculated at each unloading
point performed during the testing. Before describing the calculation of J it is
appropriate to define (redefine) the variables being used.
P - Applied load (direct measurement);
v - Load line displacement measure by COD gauge (direct measurement);
a - Crack length;
A - Area under the force-load line displacement curve;
W - Width of specimen (constant);
b - Ligament length, distance between crack front and the back of the
specimen (b = W-a;)
K - Stress intensity factor;
C
c
- Load line compliance corrected for rotation (v/p);
- Poissons ratio (constant);
E - Elastic modulus (constant);
B - Thickness of specimen (constant);

215
B
N
- Net thickness between side grooves (constant);
B
e
- Effective thickness (constant);
- Dimensionless constant;
- Dimensionless constant;
H - Half the initial distance between loading pin hole centers (constant);
R - Radius of rotation of crack centerline ((W+a)/2);
D - Half the initial distance between the knife edges (constant);
- Angle of rotation about the unbroken midsection line;

Y
- Effective yield strength .
It is convenient to separate J into elastic and plastic components:

) ( ) ( i pl i el i
J J J + = (6.2)
The elastic component is equivalent to the Griffith energy release rate assuming
plane-stress conditions:

( )
) (
2 2
1
i pl
i
i
J
E
K
J +

=
u
(6.3)
The stress concentration factor for a (C)T specimen in equation 6.3 is given by:

( )
|
.
|

\
|
=
W
a
f
W BB
P
K
i
N
i
i 2 1
(6.4)
where:

|
|
|
|
|
.
|

\
|
|
.
|

\
|

|
.
|

\
|

|
.
|

\
|
+
|
.
|

\
|

|
.
|

\
|
+
|
.
|

\
|
+ =
|
.
|

\
|
2 3
4 3 2
1
6 . 5 72 . 14 32 . 13 64 . 4 886 . 0
2
W
a
W
a
W
a
W
a
W
a
W
a
W
a
f
i
i i i i
i i
(6.5)

216
Note that the K coefficients correspond to those in Table 6-2. The plastic
component of J is essentially the plastic work required to propagate the crack
(create a new fractured surface area) which is given by:

( )
( )
( )
( ) ( )
( )
( ) ( )
( ) (
(

|
|
.
|

\
|

(
(


|
|
.
|

\
|
+ =

1
1
1
1
1
1
) 1 ) (
1
i
i i
i
N
i pl i pl
i
i
i pl i pl
b
a a
B
A A
b
J J
q
(6.6)
for a (C)T specimen where:

( )
( )
W
b
i
i
1
1
522 . 0
2

+ = q (6.7)

( )
( )
W
b
i
i
1
1
76 . 0
1

+ = (6.8)
A slight complication is that the area under the COD-load curve given by:

( ) ( )
( )
( ) ( )
( )
2
1 1
1

+
+ =
i pl i pl i i
i pl i pl
v v P P
A A
(6.9)
has to be corrected for the elastic displacement in the component. The elastic
displacement can then be subtracted out of the clip gauge data as a function of
load in the following manner:

( ) ( ) ( ) ( ) i c i i i pl
C P v v = (6.9)
The compliance of the configuration is given by:

( )
(


=
i i i i
i
i c
R
D
R
H
C
C
u u u u cos sin cos sin
(6.10)
where the angle of rotation is:

( )
|
.
|

\
|

(
(
(
(

+
|
.
|

\
|
+
=

R
D
R D
D
d
m
1
2 1
2 2
1
tan
2
sin u (6.11)

217
The crack size is given by:
5 4
3 2
677 . 650 335 . 464
043 . 106 242 . 11 06319 . 4 000196 . 1
u u
u u u
W
a
i
+
+ =
(6.12)
where:

( ) 1
1
2 1
+
=
ci e
EC B
u (6.13)
and with the effective thickness being:

( )
B
B B
B B
N
e
2

= (6.14)
Note that the compliance coefficients in equation 6.12 correspond to those in
Table 6-2. Using equations 6.3 through 6.14 to solve equation 6.2 allows for
series of J-a values to be determined at each unloading point.
i) Construction of J-R Curve: The construction of the J-R Curve was achieved by
plotting the J-a pairs calculated in the previous section. Values of J exceeding:

10 10
max max
Y Y o
B
J or
b
J
o o
= = (6.15)
were excluded, where
Y
is termed the effect yield strength and is equal to the
average of the material yield strength (
YS
) and tensile (
TS
) strength:

( )
2
TS YS
Y
o o
o
+
= (6.16)
Crack extension values exceeding:

o
b a 25 . 0
max
= A (6.17)
were also excluded from the J-R Curve.

218
j) Determination of J
IC
: The interim critical J value (J
Q
) was determined using a
graphical construct, as shown in Figure 6-10. A construction line with a slope of
M
Y
, with M=2, is plotted. Two additional parallel exclusions lines are then
plotted intersecting the abscissa at 0.15mm and 1.5mm. The data that falls
between the two parallel exclusions lines was fitted to a power law regression
line. The intersection of the regression line and a forth parallel construction line
interesting the abscissa at 0.2mm is the interim J
Q.



Figure 6-10: Graphical construction depicting the determination of J
Q
[139].


219
In order for J
Q
to be qualified as a valid interim critical J value, the following
requirements must be met:
- At least eight points must lie between the 0.15mm and 1.5mm exclusion
lines;
- At least one point must lie between the 0.15mm exclusion line and a
parallel 0.5mm construction line;
- At least one point must lie between the 0.5mm construction line and the
parallel 1.5mm exclusion line;
- The correlation coefficient of the power law regression line must be 0.96
or better;
- The physically measured initial pre-crack length and the compliance
calculated initial crack length must differ by less then 0.5mm or 0.01W,
which ever is larger;
- There must be at least three data points that lie between 0.4 J
Q
and J
Q.

Once it was established that the J
Q
values were valid, it needed to be determined
if the J value qualified as a geometry independent critical J
Q IC
. If the following
requirements are met than J =J :
Q IC

y
a y
Q
o
y
Q
Q
da
dJ
J
b
J
B o
o o
<
|
.
|

\
|
> >
A
,
25
,
25
(6.18)
All the J
Q
values measured in this study were determined to be valid and qualified
as geometry independent J
IC
values.

It is possible to use a valid J
IC
value to calculate a critical stress intensity factor in
a similar manner as done for the Griffith energy release rate where:

220
IC IC JIC
G E J E K ' ' = = (6.19)
The J-R curve holds more information than just the critical nonlinear energy
release rate. It characterizes the crack extension process. Paris [147, 148] defined a
parameter termed the tearing modulus:

|
|
.
|

\
|
=
2
Y
E
da
dJ
T
o
(6.20)
where the slope of a linear curve fit between the 0.15mm and 1.5mm exclusion lines is
multiplied by the additional terms in order to make it unitless.

6.2.3 Fractographic Analysis

Elemental analysis was performed on the fractured surfaces using an Oxford
Instruments EDS. Full field of view compositional measurements were taken on the
fractured surfaces at 100x. A JOEL JSM-6460LV SEM at working distance of 10mm,
spot size of 40, and accelerating voltage of 20KV was utilized. Approximate dimple
spacing was determined using the intercept method on fractographs at 2000x.









221
6.3 Experimental Results

6.3.1 Fracture Toughness Data

The critical nonlinear energy release rate (J
IC
) decreases with increasing yield
strengths for both the aged and SESH material, as shown in Table 6-3. The aged material
exhibited a larger scatter in toughness compared to the SESH material, as shown by the
error bars in Figure 6-11. As the aged material evolves from the peak aged condition to
overaged, the yield strength decreases and the toughness increases, however not all of the
toughness is recovered. The overaged material only recovered ~10% of the toughness
that would be expected of an underaged material at an equivalent yield strength. The rate
of decrease in toughness with increasing yield strength is twice as rapid in the aged
material compared to the SESH material. When the critical stress intensity values are
calculated using the J
IC
values the same trends are observed, as shown in Figure 6-12.

Table 6-3: Average toughness and tensile values for the aged and SESH material.
J
IC
(kJ/m
2
)
K
JIC
(MPam
1/2
)
Tearing
Modulus
Y
i
e
l
d

S
t
r
e
n
g
t
h

(
M
P
a
)
U
T
S

(
M
P
a
)
E
l
o
n
g
a
t
i
o
n

(
%
)
R
e
d
u
c
t
i
o
n

i
n

A
r
e
a

(
%
)
A
r
e
a

U
n
d
e
r

C
u
r
v
e

(
J
/
m
3
)
T4 0 61 69 16.8 313 481 20% 42% 122.4
UA 0 34 52 7.6 360 494 17% 40% 107.0
T6 0 23 42 2.9 400 493 11% 33% 70.8
OA 0 26 46 6.7 333 423 13% 44% 65.5
T4 0.16 20 39 2.8 463 537 12% 38% 82.9
W 0.17 22 41 2.7 442 519 12% 40% 81.6
W 0.31 18 38 1.7 465 540 12% 36% 88.7
T4 0.41 13 32 0.9 500 566 9% 39% 62.6
T
e
m
p
e
r
C
R

S
t
r
a
i
n
Average Toughness Values Average Tensile Values



222
0
10
20
30
40
50
60
70
80
300 350 400 450 500 550 600
Yield Strength (MPa)
J
I
C

(
k
J
/
m
2
)
0
50
100
150
200
250
300
350
400
450
43.5 48.5 53.5 58.5 63.5 68.5 73.5 78.5 83.5
Yield Strength (ksi)
J
I
C

(
l
b
f
-
i
n
/
i
n
2
)
T6 to OA
T4 to T6
CR - W
CR - T4
Linear (CR Data)

Figure 6-11: Critical nonlinear energy release rate plotted against yield strength for the
aged and SESH material.
0
10
20
30
40
50
60
70
80
300 350 400 450 500 550 600
Yield Strength (MPa)
K
J
I
C

(
M
P
a

m
1
/
2
)
0
10
20
30
40
50
60
70
43.5 48.5 53.5 58.5 63.5 68.5 73.5 78.5 83.5
Yield Strength (ksi)
K
J
I
C

(
k
s
i

i
n
1
/
2
)
T6 to OA
T4 to T6
CR - W
CR - T4
Linear (CR Data)

Figure 6-12: Fracture toughness calculated from critical nonlinear energy release rate
plotted against yield strength.

223
If the entire J-R curve is examined rather than just the critical values it is seen that
as the strength increases the curve becomes flatter, as shown in Figure 6-13. The
steepness of the J-R curves are characterized by the tearing modulus. The tearing
modulus decreases as the strength increases, as shown in Figure 6-14.

0
20
40
60
80
100
120
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4 1.5 1.6 1.7
Crack Extension (mm)
J

(
K
J
/
m
2
)
0
100
200
300
400
500
600
0.00 0.01 0.02 0.03 0.04 0.05 0.06
Crack Extension (in)
J

(
l
b
f
-
i
n
/
i
n
2
)
T4 UA
T6 OA
T4, E=0.16 W, E=0.17
W, E=0.31 T4, E=0.41
JIC

Figure 6-13: Crack extension plotted against energy release rate for both the aged and
SESH material.

224
0
2
4
6
8
10
12
14
16
18
20
22
24
300 350 400 450 500 550 600
Yield Strength (MPa)
T
e
a
r
i
n
g

M
o
d
u
l
u
s

-

M
44 49 54 59 64 69 74 79 84
Yield Strength (ksi)
T6 to OA
T4 to T6
CR - W
CR - T4
Linear (CR Data)

Figure 6-14: Tearing modulus plotted against yield strength for both the aged and SESH
material.


The side-grooving of the specimen was successful at avoiding the formation of
shear lips and prevented crack tunneling. Preliminary testing performed with nonside-
grooved C(T) specimen exhibited extensive crack tunneling and produced significant
shear lips.

6.3.2 Fractography

The fractured surfaces of the aged specimen showed transgranular and/or
intergranular fracture depending on the heat treatment they received. The material in the
T4-temper exhibited a macroscopically ductile fracture. At higher magnification the
material in the T4-temper showed transgranular dimpled rupture. Upon aging, the failure

225
mode changes from transgranular dimpled rupture to intergranular GBDF. The overaged
material showed characteristics of both failure modes, as shown in Table 6-4.

Table 6-4: Fractography observations for each processing condition.
Macroscopic (100x)
Fracture Appearance
Fracture Path Failure Mode
T
e
m
p
e
r

C
R

S
t
r
a
i
n

Ductile Brittle Transgranular Intergranular
Dimpled
Rupture
GBDF
T4 0
X
X X
UA 0 X X X
T6 0 X X X
OA 0 X X X X X
T4 0.16 X X X
W 0.17 X X X
W 0.31 X X X
T4 0.41 X X X


The fractured surfaces of the SESH material all showed intergranular dimpled
rupture, regardless of the CR strain. The fractured surfaces of the T6-temper and the
material SESH to a true strain of 0.17 in the W-tempter are shown in Figure 6-15
through Figure 6-18 for a comparison of the different fracture modes. The SEM
fractographs for all eight processing conditions are shown in Appendix F. The dimples
were more closely spaced on the fractured surfaces of the aged material compared to the
SESH, as shown in Figure 6-19. It should be recognized that the dimple spacings are
approximate values being that they were determined using linear measurements on non-
planer surfaces.

226

Figure 6-15: SEM fractograph looking at the RD plane of the material in the T6-temper at
100x showing a brittle appearance.




Figure 6-16: SEM fractograph looking at the RD plane of the material in the T4-temper at
1,000x showing GBDF.

227

Figure 6-17: SEM fractograph looking at the RD plane of the material SESH in the
W-temper to a true strain of 0.17 at 100x showing a ductile appearance.




Figure 6-18: SEM fractograph looking at the RD plane of the material SESH in the
W-temper to a true strain of 0.17 at 1,000x showing intergranular dimpled rupture.

228
y = 20.533x + 4.7383
y = 26.566x + 21.397
0
10
20
30
40
50
60
70
80
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Approximate Dimple Spacing (um)
K
J
I
C

(
M
P
a

m
1
/
2
)
0
10
20
30
40
50
60
70
0 10 20 30 40 50 60 70
Approximate Dimple Spacing (in E-6)
K
J
I
C

(
k
s
i

i
n
1
/
2
)
T4 to T6
OA
CR-W
CR-T4

Figure 6-19: Approximate dimple spacing plotted against fracture toughness for each
processing condition.


There was no measurable difference in the elemental composition of the fractured
surfaces regardless of the processing route used to achieve the properties. Of the alloying
elements present in the material, Cu was most prevalent on the fractured surface followed
by Mg, and Mn, as shown in Figure 6-20. The average content level of alloying elements
present on the fractured surface is ~60% higher (based on the % Al measured) than when
measured on a random cross-section, as shown in Table 6-5. The Cu, Mn, and Fe level
measured on the fractured surface exceed the composition requirements for AA2524.


229
0.0
1.0
2.0
3.0
4.0
5.0
6.0
7.0
W
e
i
g
h
t

%
T4 OA UA T6 Ave. W,
E=0.17
T4,
E=0.16
W,
E=0.31
T4,
E=41
Copper (Cu)
Magnesium (Mg)
Manganese (Mn)
Iron (Fe)
Figure 6-20: Elemental compostion of the fratured surfaces for each processing condition.


Table 6-5: Average elemental composition of the fractured surfaces, bulk material, and
the range specified for AA2524.
Average Values
Element (wt%)
Fractured
Surfaces
Polished
Surfaces
AA2524
Aluminum (Al) 90.7 94.2 92.5 - 94.4 %
Copper (Cu) 6.4 4.0 4.00 - 4.50 %
Magnesium (Mg) 1.2 1.3 1.20 - 1.60 %
Manganese (Mn) 1.0 0.5 0.450 - 0.700 %
Zinc (Zn) 0.1 0.1 0.150 %
Titanium (Ti) < 0.1* < 0.1* 0.100 %
Iron (Fe) 0.5 0.1 0.120 %
Silicon (Si) < 0.1* < 0.1* 0.0600 %
Chromium (Cr) < 0.1* < 0.1* 0.0500 %
* Outside the resolution of the equipment.




230
6.4 Discussion

6.4.1 Fracture Toughness

The fracture toughness of both the aged and SESH material was characterized
using the critical nonlinear energy release rate rather than the critical stress intensity
value due to geometric constraints. Using the values in Table 6-3, it is estimated that a
C(T) specimen thickness of 121mm would be required to obtain valid K
IC
values for the
material in the T4-temper and 22mm for the toughest SESH material. The
equipment/loads required to test specimen of this size would not have been a issue if it
were possible to acquire material of these thicknesses. It should be noted that the critical
stress intensity value (K
JIC
) calculated using the critical J-integral value (J
IC
) is
equivalent to the toughness value obtained through plane-strain fracture conditions (K
IC
)
only if all of the experimental and dimensional post test criteria are satisfied. The relative
agreement of these values is also dependent on the slope of the R-curve. The flatter the R-
curve the closer the critical values. This is because K
IC
is measured at 2% percent crack
extension (relative to the ligament length), while K
JIC
is measured from an absolute crack
extension (0.2mm). All criteria to qualify K
JIC
as a valid geometry independent
measurement of toughness were satisfied in the current work.
The R-curves for all of the processing conditions exhibited an initial region of
crack tip blunting followed by crack extension. The lower strength aged conditions (T4,
UA, OA) absorbed more energy per unit crack extension resulting in steeper R-curves
with tearing modulus values ranging form two to over five times higher than the SESH
material. The steeper R-curves are most likely the result of the formation of large plastic

231
zones and possibly the loss of stress triaxiality after initial crack extension. The R-curves
for the T-6 temper and the SESH material are all relatively flat with tearing modulus of
less than 3. None of the R-curves reached steady state crack growth which is indicative of
a fully developed plastic zone. Although all of the specimens tested, regardless of the
processing, produced increasing R-curves the requirements were met to obtain valid
geometry independent J
IC
values. It should be noted that while the K
JIC
values
determined from the early stages of the R-curves are geometry independent the later
portion of the R-curves most likely are not. The full R-curves and tearing modulus are
dependent on the specimen geometry.
It was shown in the previous Chapter that the yield strength values of the aged
AA2524 were within the expected ranged for 2xxx series alloys. It was also established
that SESH of AA2524 could be used to produce yield strengths exceeding those of the
T6-temper and equivalent to the higher strengths reported in other 2xxx series
alloys/conditions. The superior tensile properties achieved in this alloy are of limited
application if the damage tolerance of the material is negatively impacted. The SESH
process produced material with an equivalent toughness to the baseline AA2524 material
in the T6-temper but with an 11% higher yield strength. If that same data point (CR-W,
true stain of 0.17) is compared to published strength-toughness data for 2x24 [132] alloys
at an equivalent yield strength the fracture toughness is ~65% higher, as shown in Figure
6-21. The AA2524 material was specifically developed to have an improved fracture
toughness over the original 2024 alloy. However, it is reported that the improvement in
fracture toughness relative to 2024 is 15-20% [140]. The lower levels of impurities in
AA2524 does not account for the ~65% improvement (or ~40% improvement when

232
closest toughness values are compared at equivalent yield strengths). The AA2524 T6
and overaged material are within this 15-20% improvement range, as would be expected.
Aluminum alloys are never used in the underaged condition and AA2524 is very rarely
used in the T4 condition due to general corrosion and stress corrosion cracking (SCC)
susceptibility. If 2x24 alloys are used in the T4-temper they are often restricted to thin
sheet gauge thicknesses, due to the quench sensitivity of the material, which does not
facilitate standard fracture toughness testing. As a result, published toughness values for
the underaged and T4 condition are not readily available or are usually not geometry
independent.

0
10
20
30
40
50
60
70
80
300 350 400 450 500 550 600
Yield Strength (MPa)
K
I
C
/
J
I
C

(
M
P
a

m
1
/
2
)
0
10
20
30
40
50
60
70
43.5 48.5 53.5 58.5 63.5 68.5 73.5 78.5 83.5
Yield Strength (ksi)
K
I
C
/
J
I
C

(
k
s
i

i
n
1
/
2
)
2xxx
2x24
2524 (T6 to OA)
2524 (T4 to T6)
2524 (CR - W)
2524 (CR - T4)

Figure 6-21: Fracture toughness plotted against yield strength for 2524 and 2xxx series
alloys.



233
6.4.2 Fractography

The improved fracture toughness in the SESH material is a result of the failure
mode changing from GBDF to intergranular dimpled rupture, which is a consequence of
the processing route and subsequent microstructure. Typically, precipitation hardened
aluminum alloys fail by GBDF which is a result of the formation of GB precipitates
during quenching and subsequent formation/growth during artificial aging. Given that
grain boundaries are a three dimensional interconnected network throughout the
microstructure they create a preferential fracture path when decorated with precipitates.
The fractographs of the aged material did in fact show that the fracture was intergranular
with a smaller dimple spacing than the SESH material. This preferential GB fracture path
is not present (or at least not dominant) in the SESH material, making strain localization
between dimple sites more difficult and resulting in greater toughness values.
The solute levels on the fractured surfaces were ~60% greater than on a random
cross section. The higher solute levels are a result of crack initiation/growth preferentially
occurring along dispersoids, precipitates, and constituents within the microstructure. The
overall elemental composition of the aged and SESH fractured surfaces showed no
discernable difference, although the fracture morphology of the aged and SESH material
was substantially different. This observation supports the expected result that the fracture
process is more dependent on the spatial distribution of the solute (precipitated or in
solution) than its absolute value. Any composition differences that may have existed were
not within the resolution of the equipment. Elemental measurements with greater
resolution would have been expected to show local compositions biased toward specific

234
phases present at the GBs in the aged material and within the grains for the SESH
material.
The SESH material may also be benefiting from the absence of a precipitate free
zone (PFZ). It is known that PFZ form adjacent to the GBs in 2x24 alloys during artificial
aging [141, 142]. What is unclear is how the process of void formation, growth, and
coalescence , differs around a precipitate within a grain verse a GB precipitate in a PFZ.
Studies have shown that strain accumulates more rapidly in the PFZ than in the matrix of
Al alloys [12, 144-146]. The localized accumulation of strain in the PFZ of the aged
material may promote void growth and coalescence, negatively influencing the fracture
toughness of the material. The presence of the PFZ may also explain why the elemental
composition of the aged and SESH fractured surfaces were indistinguishable. The
fracture of the aged material follows the solute rich GBP; however, the fracture is also
preferentially traveling through the solute depleted PFZ.

6.5 Conclusions

SESH is a viable processing route for producing improved combinations of
strength and toughness in AA2524. At equivalent yield strengths, the SESH
material exhibited an increase in toughness ranging from ~40% to 60% of typical
2x24 values. At an equivalent toughness to that of the T6 material, the SESH had
an 11% higher yield strength.
The SESH material exhibited transgranular dimpled rupture in all processing
conditions as opposed to the GBDF seen in the aged material. The dimple size

235
and spacing was also larger in the SESH material. The SESH material may have
benefited from the absence of a PFZ.

6.6 Future Work

The natural extension of the current project would be the investigation of the
fatigue properties of the SESH material exhibiting improved combinations of strength
and toughness. Fatigue performance is a critical design parameter in many applications.
There is a possibility that cyclic loading may result in a partial recovery of the dislocation
structure, resulting in the softening of the material. There is also a possibility that the
absence of a PFZ and lower frequency of particles/precipitates may produce improved
fatigue properties. These, and other questions regarding the fatigue performance of the
SESH microstructures, need to be investigated.
There is a potential to further improve the properties of SESH aluminum by
performing a thermal recovery after processing. Preliminary work performed by [5]
showed that a thermal recovery could significantly improve tensile ductility with only a
moderate loss in yield strength, provided that precipitation does not occur. This increase
in ductility would be expected to also increase the fracture toughness of the material. The
more stable microstructure present after the thermal recovery many also positively effect
fatigue properties.
A derivative of aluminum alloy 2024 was selected for the current studies because
it exhibited the greatest potential to effectively SESH. As shown in the current work, the
ability to achieve strengths equivalent to or higher than those produced by precipitation
hardening does not require exceptionally large strains. The best properties may be

236
achieved in the less effectively SESH 7xxx series alloys. These more modern, less
quench sensitive alloys are available in thicker sections to facilitate processing/testing
and exhibit higher T4 properties. Moving forward, it may be a better option to use the
knowledge gained in the current study to explore the 7xxx series alloys rather than further
characterizing the 2xxx series.



















237
CHAPTER 7. PROJECT SUMMARY

The ultimate objective of this project was to determine if improved combinations
of yield strength and fracture toughness could be achieved in aluminum alloys through
SESH processing. The data compiled within this document shows that improved
combinations of strength and toughness can in fact be achieved in AA2524. The
experimental progression used to make this final determination has been separated into
three core chapters, each of which contains its own conclusions and future work section.
The first of these core chapters is entitled "Effect of GP Zone Formation on Strain
Hardening Rate", (Chapter 4). This chapter addresses the upfront processing challenges
encountered during the project, such as the mid-plane cracking, and establishes the
practical processing limitations. The primary conclusion of this chapter is that: as GP
zone formation occurs in AA2524, the SHR diminishes and SESH becomes less effective.
The second core chapter is entitled "Improved Strength Through Solute Enhanced
Strain Hardening" (Chapter 5). The baseline tensile properties are developed in these
sections and the through thickness uniformity of the SESH is addressed. The primary
conclusion of this chapter is that: SESH is capable of achieving yield strengths equivalent
to or greater than those produced by precipitation hardening. At an equivalent yield
strength the overall tensile properties (strength and ductility) of the SESH material were
improved.
The final core chapter is entitled "Effect of Solute Enhanced Strain Hardening on
Toughness" (Chapter 6). This chapter builds off the previous two and addresses the
different failure modes observed for the various processing conditions, showing that the
aged material failed by GBDF and the SESH material by intergranular dimpled rupture.

238
The primary conclusion of this chapter is that: SESH is a viable processing route for
producing improved combinations of strength and toughness in AA2524.


239
LIST OF REFERENCES




1. Brush Wellman Inc., MATERIAL SAFETY DATA SHEET - Copper Beryllium
Wrought Alloy, www.brushwellman.com.

2. InterCorr International, Corrosion Source, Corrosion Source Handbook,
www.corrosionsource.com.

3. J.E. Dorn, P. Pietrosdky, and T.E. Tietz, Trans ASM 188, 933 (1950).

4. C.G. Schmidt and A.K. Miller, Acta Metall., 30, 615 (1982).

5. R.D.Doherty and J.McBride (1993) In Aluminum Alloys for Packaging Eds.
J.G.Morris, H.D. Merchant, E.J. Westerman and P.L. Morris, TMS-AIME
Warrendale PA. 15086, pp 347-368.

6. R.D. Doherty and S.F. Baumann (1993) In Aluminum Alloys for Packaging Eds. J.G.
Morris, H.D. Merchant, E.J. Westerman and P.L. Morris, TMS-AIME Warrendale
PA. 15086, pp 369-391

7. E.M. Shaji, S.R. Kalidindi, R.D. Doherty, and A.S. Sedmak, Mater. Sci. and Eng.
(2003) A340 163-169.

8. J. Frenkel, Z. Phys., vol. 37, p. 572, (1926).

9. T.H. Courtney, ISBN:0-07-013265-8, p.83, McGraw-Hill (1990).

10. G.I. Taylor, Proc. R. Soc. London, vol. 145A, p362, (1934).

11. E. Orowan, Z. Phy., vol.89, pp.605, 614, 634, (1934).

12. M. Polanyi, Z. Phys., vol. 89, p. 660 (1934).

13. F.R. Nabarro, Theory of Crystal Dislocations, Oxford: Clarendon Press (1967).

14. G.E. Dieter, Mechanical Metallurgy, McGraw-Hill (1961).

15. I. Schnell, Ph.D. Thesis, University of Bremen, Germany (2002).

16. H. Wiedersich, J. Metals, 16, p.425 (1964).

17. D. Kuhlmann-Wilsdorf, Trans. Met. Soc. AIME 224, 1047 (1962).


240
18. D. Kuhlmann-Wilsdorf, Work Hardening, J.P. Hirth and J. Weertman, Eds.,Gordon
&Breach, New York, p. 97, (1968).

19. D. Kuhlmann-Wilsdorf, Work Hardening in Tension and Fatigue, 4
th
ed., A. W.
Thompson, Ed., AIME, New York, p. 1, (1977).

20. R.W. Hertzberg, Deformation and Fracture Mechanics of Engineering Materials,
John Wiley & Sons, New York (1996).

21. A. Seeger, Dislocations and Mechanical Properties of Crystals, p. 243, John Wiley
& Sons, New York, (1957).

22. G.I. Taylor, J. Inst. Met. 62, 307 (1938).

23. E. Schmid, Z. Elektrochem., vol. 37, 447 (1931).

24. R. VonMises, Z. Angew. Math. Mech., vol. 8, p. 161 (1928).

25. M.F. Ashby, Phil. Mag., ser. 8, vol. 21, pp. 399-424, (1970).

26. W. Boas and M.E. Hargreaves, Proc. R. Soc. London Ser .A, vol. A193, p. 89
(1948).

27. E.O. Hall, Proc. Phys. Soc. London, vol. 643, p. 747, (1951).

28. N.J. Petch, J. Iron Steel Inst. London, Vol. 173, p. 25, (1953).

29. F.C. Frank and W.T. Read, Phys. Rev., vol.97, pp. 722-723 (1950).

30. J.D. Eshelby, F.C. Frank, and F.R.N. Nabarro, Plil. Mag. 42, 351 (1951).

31. M.A. Meyers and K.K. Chawla, Mechanical Behavior of Materials, p. 271, Prentice
Hall, New Jersey, (2002).

32. R.W. Armstrong, Advances in Materials Research, Vol. 5, R.F. Bunshah, p 101,
Wiley-Interscience (1971).

33. Fan, G.J., et al., Plastic deformation and fracture of ultrafine-grained Al-Mg alloys
with a bimodal grain size distribution. Acta Materialia, 2006. 54(7): p. 1759-1766.

34. Fu, H.-H., D.J. Benson, and M.A. Meyers, Analytical and computational
description of effect of grain size on yield stress of metals. Acta Materialia, 2001.
49(13): p. 2567-2582.


241
35. Furukawa, M., et al., Microhardness measurements and the Hall-Petch relationship
in an Al-Mg alloy with submicrometer grain size. Acta Materialia, 1996. 44(11): p.
4619-4629.

36. Meyers, M.A., D.J. Benson, and H.-H. Fu, Grain-size-yield stress relationship:
analysis and computation. Advanced Materials for the 21st Century, 1999: p. 499-
512.

37. Meyers, M.A., A. Mishra, and D.J. Benson, Mechanical properties of
nanocrystalline materials. Progress in Materials Science, 2006. 51: p. 427556.

38. Carlton, C.E. and P.J. Ferreira, What is behind the inverse Hall-Petch effect in
nanocrystalline materials? Acta Materialia, 2007. 55(11): p. 3749-3756.
39. A.H. Cottrell, Trans-AIME, 212, 192 (1958).

40. Meyers, M.A. and E. Ashworth, A model for the effect of grain size on the yield
stress of metals. Philosophical Magazine A, 1982. 46(5): p. 737-759.

41. Callister, W.D., Materials Science and Engineering: An Introduction. 6th ed. 2003,
Hoboken: John Wiley & Sons, Inc. 820.

42. Cottrell, A.H., Dislocations and plastic flow in crystals, Oxford: Clarendon Press
(1953).

43. P. Haasen, Krist. U. Techn. 2, 251 (1967).

44. P. Haasen, Proc. Int. Conf. Tokyo, Trans. Japan. Inst. Met. Suppl. 9, XL (1986).

45. W.E. Nixon and J.W. Mitchell, Proc. Roy. Soc. A376, 343 (1981).

46. R. Labusch, Phys. Stat. Sol 41, 659 (1970).

47. R. Labusch, Acta Metall. 20,917 (1972).

48. P. Haasen, Mechanical Properties of Solid Solutions, R.W. Cahn and P Haasen,
eds., vol. 3, Chapter 23, (1996).

49. C.S. Smith, Trans. Metall. Soc. AIME 175, 15 (1948).

50. G. Couturier et al., Materials Science Forum, 467-470, 1009 (2004).

51. Servi and Turnbull, Acta Met, Vol. 14, (1966).

52. D.A. Porter and K.E. Easterling, Phase Transformations in Metals, CRC Press, UK
(1992).


242
53. W. Ostwald, Z. Phys. Chem., 34, 495, (1900).

54. E. Orowan, Symposium on Internal Stresses, p. 451, Institute of Metals, London
(1947).

55. A. Kelly and R.B. Nicholson, Prog. Matls. Sc., 10, 151 (1963).

56. V. Gerold and H.P. Karnthaler, Acta Met., 37, 2177 (1989).

57. Courtesy of M.V. Heimendakl, pp. 485 of [31].

58. L.T. Anderson, Fracture Mechanics, 3
rd
ed., CRC Press, Boca Raton, Fl (2005).

59. ASTM International, West Conshohocken, Pennsylvania, USA.

60. A.A. Griffith, Philos. Trans. R. Soc. London, vol. 221A, pp. 163-198 (1920).

61. C.E. Inglis, Trans Inst. Nav. Archit., vol. 55, pt. I, pp. 219-230 (1913).

62. E. Orowan, Fracture and Strength in Solids, Reports on Prog. in Phys., Vol. XIIp.
185 (1948).

63. G.R. Irwin, Fracture Dynamics, American Soc. of Met, Cleveland, OH, pp. 147-166
(1948).

64. G.R. Irwin, Sagamore Research Conference Proceedings, Vol. 2, pp. 289-305
(1956).

65. Y. Murakami, Stress Intensity Factors Handbook, 3
rd
ed. (1999).

66. Xue-Ren Wu and A. Janne Carlsson, Weight Functions and Stress Intensity Factor
Solutions (1991).

67. H.M. Westergaard, Trans, ASME, J. Appl. Mech. 61, 49 (1939).

68. G.R. Irwin, Handbuch der Physik, vol. 6 Springer, Berlin, p. 551 (1958).

69. Barsom and Rolfe, Fracture and Fatigue Control in Structures, 2dn ed., Prenttice-
Hall, Englewood Cliffs, NJ (1987).

70. J.R. Rice, Journal of Applied Mathematics, Vol. 35, pp. 379-386 (1968).

71. J.R. Rice and E.P. Sorenson, J. Mech. Phys. Soilds 26, 163 (1978).

72. J.R. Rice, Treatise on Fracture, vol. 2, H. Liebowitz, Ed., Academic, New York, p.
191 (1968).

243

73. J.D. Eshelby, Phil. Trans. Roy. Soc. London, A244, 87 (1951).

74. NMAB Report 328, Rapid Inexpensive Tests for Determining Fracture Toughness
(1976).

75. G.R. Irwin, Sagamore Research Conference Proceedings, vol. 4, Syracuse
University Research Institute, Syracuse, NY, pp. 63-78 (1961).

76. A.K. Vasudevan and R.D. Doherty, Overview #58. Acta Metall., 35 1193-1219.
(1986).

77. R. Seng and E. Spuhler, Metal Progress, March 1975, copyright American Society
for Metals.

78. ASTM E399, ASTM International, West Conshohocken, Pennsylvania, USA
(2009).

79. D. Broek, Eng. Fract. Mech. 5, 55 (1973).

80. J.T. Staley, Effect of Quench Path on Properties of Aluminum Alloys, p. 396, Alcoa
Laboratories, Pennsylvania, USA.

81. J.T. Staley, R.D. Doherty, and A.P. Jaworski, Met. Trans., Vol. 24A, p. 2417
(1993).

82. Alcoa Green Letter, Vol. II, Alcoa 7085 Hand Forgings, 1
st
. Ed., July (2006).

83. P.T. Unwin and G.C. Smith, J. Inst. Metals 97, 299 (1969).

84. J.C. McMahon Jr, V. Vitek and J. Kamada, Development in Fracture Mechanics,
2
nd
ed., p. 193. Appl. Sci., London (1981).

85. J.D. Embury and E. Nes, Z. Metallk. 65, 45 (1974).

86. I. Kirman, Metall. Trans. 2, 1761 (1971).

87. R.S. Sanders, S.F. Baumann, and H.C. Stumpf, Aluminum Alloys Contemporary
Research and Applications, Eds. A.K. Vasudevan and R.D. Doherty, Academic
Press, San Diego, CA, pp. 66-105.

88. Airbus, Innovation and Technology. www.airbus.com

89. Automotive Aluminum Association, Inc., Environmental Advantages
http://www.autoaluminum.org


244
90. Material Property Data, Metals. www.matweb.com

91. International Energy Agency. www.iea.org

92. R. D. Doherty, J. Liu and R. E. Sanders Jr.(1993) Aluminum Alloy Sheet for Food
and Beverage Containers Patent # 5,192,378 Assigned to Alcoa Issued on March 9,
1993.

93. E. Hashimoto, K. Ono, and T. Kino, Efficiency of Dislocation as a Sink for
Vacancies in Deformed Aluminum, J. Phy Soc Japan, vol. 42, No.3 (1977).

94. D. Turnbull, H.S. Rosenbaum, and H.N. Treaftis, Kinetics of Clustering in some
Aluminum Alloys, Acta Met, vol. 8 (May, 1960).

95. D. Turnbull and R.L. Cormia, Kinetics of Later Stages of Clustering in Al-Cu
Alloy, Acta Met, vol. 8 (Nov., 1960).

96. L.A. Girifalco and H. Herman, A Model for the Growth of Guiner-Preston Zones
The Vacancy Pump, Acta Met, vol. 13 (1965).

97. M.B. Prime and M.R. Hill, Residual Stress, Stress Relief, and Inhomogeneity in
Aluminum Plate, Scripta Mat., Vol. 46, No. 1, pp. 77-82 (2002).

98. J.S. Robinson and D.A. Tanner, Mat. Sci. Forum, vols. 404-407 pp. 355-360
(2002).

99. D.A. Tanner and J.S. Robinson, Mat. Sci. and Tech., Vo. 22, No. 1, p.77 (2006).

100. M. Koc, J. Culp, and T. Altan. J. Mat. Pros. And Tech., 174, pp342-354 (2006).

101. G.E. Totten and D.S. MacKenzie, Handbook of Aluminum, Vol. 1, CRC Press
(2003).

102. T. Croucher, Heat Treating, pp. 24-26, Sept. (1980).

103. T. Croucher, Heat Treating, pp. 34-37, Oct. (1980).

104. A. Considere, Ann. Points et chausses, vol. 9, ser. 6, pp. 574-775 (1885).

105. D.V. Wilson, M. Zandrahimi, and W.T. Roberts, Acta Met., vol. 38, No. 2, pp. 215-
226 (1990).

106. M. Zandrahimi, S. Platias, D. Price, D. Barrett, P.S. Bate, W.T. Roberts, and D.V.
Wilson, Met Trans A, Vol. 20A (1989).

107. P.S. Bate, Met Trans A, Vol. 24A (1993).

245

108. U.F. Kocks and T.J. Brown, Acta Met, vol. 4, (1966).

109. AMS 2770H, Heat Treatment of Wrought Aluminum Parts, SEA International,
Warrendale, PA, (Revised 2006).

110. T. Croucher, Minimizing Machining Distortion in Aluminum Alloys Through
Successful Application of Uphill Quenching, Journal of ASTM International,
(2008).

111. T. Croucher, A Study of Different Uphill Quenching Variables on the Reduction of
Residual Stress in Aluminum Alloys, Unpublished document, (2008).

112. T. Croucher, Using Uphill Quenching to Effectively Stabilize Machined Aluminum
Parts. Tom Croucher and Associates, (2006).

113. K. Boiko, Aluminum Alloys Finds Solution for Stresses can be Uphill, Heat
Treating, (1983).

114. R.S. Barker, H.N. Hill, and L.A. Willey, Transactions of The ASM, vol. 52 (1960).

115. N. Yoshihara and Y. Hino, Residual Stress III Sci and Tech, Vol. 2, pp. 1140-
1145 (1992).

116. H.N. Hill, R.S. Barker, and L.A. Willey, Trans of American Soc. for Met., vol. 52,
pp. 1140-1145 (1960).

117. H. Chandler. Heat Treater's Guide. ASM International, Materials Park, OH, 1996.

118. ASTM E 18-08b, Test Methods for Rockwell Hardness of Metallic Materials,
ASTM International, West Conshohocken, Pennsylvania, USA (2008).

119. ASTM E 1004-02, Standard Practice for Determining Electrical Conductivity Using
the Electromagnetic (Eddy-Current) Method, ASTM International, West
Conshohocken, Pennsylvania, USA (2002).

120. ASTM E 9, Test Methods for Compression Testing of Metallic Materials at Room
Temperature, ASTM International, West Conshohocken, Pennsylvania, USA
(2009).

121. AMS 3025C, Polyalkylene Glycol Heat Treat Quenchant, SEA International,
Warrendale, PA, (Revised 2009).

122. S. MacKenzie, Heat Treating Aluminum for Aerospace Applications, Heat Treating
Progress (July, 2005).


246
123. T. Croucher, Using Glycol to Effectively Control Distortion and Residual Stresses
in Heat Treated Aluminum, J. of ASTM International (JAI), (2008).

124. T. Croucher, A Systems Approach to Quenching Aluminum aids Dimensional
Stability, Light Metals, Heat Treating, Vol. XIV, No.11, pages 18-19 (1982).

125. C.J. Hovanec and R.D. Doherty, Deformation of High Solute Aluminum Alloys and
the Inhibition of GP Zone Formation, Solid-to-Solid Phase Transformations in
Inorganic Materials, Vol. 1, Diffusional Transformations, TMS, (2005).

126. J.H. Hollomon, Trans. AIME, vol. 162. pp. 268-290 (1945).

127. U.F. Kocks, H. Mecking, Prog. Mater. Sci. 48 (2003) 171273.

128. D. Ablitzer and M. Gantois, La Diffusion dans les Milieux Condenses. Theorie et
Apllications, CEN Saclay, vol. 1, p. 299 (1976).

129. W. DeSorbo and D. Turnbull, Physical Review, Vol. 115, No. 3, (1959).

130. ASM Handbook, Mechanical Testing and Evaluation, Vol. 8, ASM International,
Materials Park, OH, (1990).

131. U.F. Kocks, M.G. Stout, and A.D. Rollett. The Influence of Texture on Strain
Hardening, Proceedings of the 8
th
International Conference on the Strengths of
Metals and Alloys, pp. 25-34 (1988).

132. J.G. Kaufman, Fracture Resistance of Aluminum Alloys, ASM International,
Materials Park, OH (2001).

133. ASTM E 92-82, Standard Test Method for Vickers Hardness of Metallic Materials,
ASTM International, West Conshohocken, Pennsylvania, USA (2003).

134. ASTM E 8-08, Standard Test Method for Tension Testing of Metallic Materials,
ASTM International, West Conshohocken, Pennsylvania, USA (2008).

135. H. Biloni and W.J. Boettinger, Physical Metallurgy 4
th
ed., vol. I, Chap. 8, Elservier
BV (1996).

136. M.C. Flemings, Solidification Processing, Mcgraw-Hill College, (1974).

137. LF. Mondolfo, Aluminum Alloys: Structure and Properties, Butterworths Pub.
(1979).

138. ASTM E 1820, Standard Test Method for Measurement of Fracture Toughness,
West Conshohocken, Pennsylvania, USA (2009).


247
139. J.A. Joyce, Elastic-Plastic Fracture: Laboratory Test Procedures, MNL 27, ASTM
International, West Conshohocken, Pennsylvania, USA, Printed in Scranton PA
(1996).

140. J.C. Williams and E.A. Starke, Progress in Structural Materials for Aerospace
Systems, Acta Mat, 51, pp. 5775-5799 (2003).

141. W. Zhang and G.S. Frankel, Electrochimica Acta, 48, pp. 1193-1210 (2003).

142. M.O. Speidel, M.V. Hyatt, in: M.G. Fontana, R.W. Staehle (Eds.), Advances in
Corrosion Science and Technology, vol. 2, Plenum Press, New York, p. 115 (1972).

143. A. Deschamps, S. Esmaeili, and W.J. Poole, and M. Militzer, J. De Phy. IV, vol. 10,
pp. 151-156 (2000).

144. M. Graf and E. Hornbogen, Acta metall, 25, 883 (1977).

145. A.J. Cornish and M.K.B. Day, J. Inst. Metals, 99, 377 (1971).

146. T.H. Sanders, E.A. Ludwiczak, and R.R. Sawtell, Mater. Sci. Engng, 43, 247
(1980).

147. P.C. Paris, ASTM STP 631, p.3 (1977).

148. P.C. Paris, H. Tada, A. Zahoor, and H. Ernst, ASTM STP 668, p. 251 (1978).




248
APPENDIX A. Compression Stress-Strain Curves




This appendix contains the compressive true stress-strain curves for each of the
three natural aging conditions (W-temper, 1/2H-temper, and T4-temper), after being
subjected to various cold rolling strains. The yield strength values are proved in Table
4-2 and as a function of prior rolling strain in Figure 4-22. The stress-strain curves in this
appendix were used to produce the SHR relationships in Figure 4-28 through Figure
4-33.

0
100
200
300
400
500
600
700
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
True Compressive Strain
T
r
u
e

C
o
m
p
r
e
s
s
i
v
e

S
t
r
e
s
s

(
M
P
a
)
0
20
40
60
80
100
T
r
u
e

C
o
m
p
r
e
s
s
i
v
e

S
t
r
e
s
s

(
k
s
i
)
W, E=0.0
W, E=0.13
W, E=0.23
W, E=0.31
W, E=0.41

Figure A-1: True stress plotted against true strain for cylindrical compression specimen
tested after being cold rolled to various strains in the W-temper.



249
0
100
200
300
400
500
600
700
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
True Compressive Strain
T
r
u
e

C
o
m
p
r
e
s
s
i
v
e

S
t
r
e
s
s

(
M
P
a
)
0
20
40
60
80
100
T
r
u
e

C
o
m
p
r
e
s
s
i
v
e

S
t
r
e
s
s

(
k
s
i
)
1/2H, E=0.0
1/2H, E=0.15
1/2H, E=0.21
1/2H, E=0.41

Figure A-2: True stress plotted against true strain for cylindrical compression specimen
tested after being cold rolled to various strains in the 1/2H-temper.
0
100
200
300
400
500
600
700
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
True Compressive Strain
T
r
u
e

C
o
m
p
r
e
s
s
i
v
e

S
t
r
e
s
s

(
M
P
a
)
0
20
40
60
80
100
T
r
u
e

C
o
m
p
r
e
s
s
i
v
e

S
t
r
e
s
s

(
k
s
i
)
T4, E=0
T4, E=0.12
T4, E=0.22
T4, E=0.31
T4, E=0.41

Figure A- 3: True stress plotted against true strain for cylindrical compression specimen
tested after being cold rolled to various strains in the T4-temper.

250
APPENDIX B. Deformation Al Alloys & the Inhibition of GP Zone Formation




TMS (The Minerals, Metals & Materials Society), 2005

Deformation of High Solute Aluminum Alloys and the Inhibition of GP Zone
Formation

Christopher J. Hovanec, Roger D. Doherty
Drexel University, Department of Materials Science and Engineering
3141 Chestnut St.; Philadelphia, PA 19104

Natural Aging, GP Zones, Aluminum, ECAP, Strain Hardening

Abstract

In order to determine the relative effectiveness of solute in solution versus it clustered
into GP zones on promoting strain hardening, a series of rolling experiments were carried
out on 2524, an Al-Cu-Mg-Mn alloy, after different periods of natural aging. The results
clearly show that as the alloy is naturally aged, forming GP zones, the strain hardening
increment falls, for a given reduction. This is accompanied by a parallel fall in electrical
conductivity, of which the magnitude decreases with the formation of GP zones, once
again indicating a reduction in dislocation retention when the solute forms clusters. It was
also found after a rolling reduction to a strain of 0.4, that the hardness rise and electrical
conductivity fall characteristic of natural aging is totally inhibited. For a smaller rolling
strain of only 0.1, there is a small residual hardness increment due to natural aging but
this increment of hardening is only 15% of that found in the un-deformed sample. It is
suggested that the loss of the natural aging response as a result of cold rolling is due to
the loss of the excess quenched-in vacancies after solution treatment. These vacancies,
needed for room temperature aging, appear to be absorbed by the dislocations introduced
by cold rolling. Since the origin of solute enhanced strain hardening in cold rolled
aluminum alloys is known to be solute inhibition of dynamic recovery, the present results
indicate that this inhibition is less effective when the solute is clustered into GP zones.

Introduction

Aluminum alloys are traditionally strengthened by precipitation hardening. This is an
effective method for increasing the yield strength, convenient for complex shapes and is
easily performed on an industrial scale. An unfortunate consequence is that less coherent
precipitates form on the grain boundaries degrading the fracture toughness of the alloy
[1]. This embrittling effect can be avoided by strain hardening. Strain hardening has been
largely ignored as a useful strengthening method for Al alloys, with the exception of the
beverage can industry [2]. The reason for this is that Al has a large stacking fault energy
which allows for dynamic recovery to easily occur, resulting in ineffective work
hardening.

251

It has been shown by Doherty et al. [3&4] that high solute Al alloys can be effectively
strain hardened provided that the solute is not precipitated. It is believed that the solute
inhibits dynamic recovery. This is a significant finding from an engineering standpoint
due to the development of novel processing techniques, such as Equal Channel Angular
Pressing (ECAP), which is capable of imposing large stains ( > 10) while retaining a
cross sectional area capable of being used in load baring applications. In the investigation
performed by [3] the deformation was imposed by cold rolling up to strains of 2.3 for a
range of alloys with increasing amounts of solute, as shown in Figure 1.

0
100
200
300
400
500
600
700
0 1 2 3 4 5 6 7
Strain
Y
i
e
l
d

S
t
r
e
s
s

(
M
P
a
)
3Cu-1Mg
2Cu-0.5Mg
1Cu-0.5Mg
Al 99.99%

Figure 1. Yield stress plotted against strain for Al alloys with increasing amounts of
solute [3]. The alloy with 3wt% Cu and 1 wt% Mg has been extrapolated from a strain of
2.3 to 6.

It can be seen that as the amount of solute increases so does the flow stress at equivalent
strains. If these curves are extrapolated out to large strains using the Hollomon-Stumpf
equation (1), yield strengths of over 600 MPa should be achievable. The Hollomon-
Stumpf equation was developed for cold rolling where is the tensile yield strength,
y R

is the imposed strain, K is the yield stress at a strain of 1, and n is the strain hardening
exponent. It has been shown that K increases with solute.

n
R y
Kc o = (1)

Unfortunately, when this work was performed there was very little attention paid to the
formation of GP zones and their effect on strain hardening. This prompts the questions;
1) is strain hardening more effective with solute in solution or precipitated as GP zones,
and 2) what is the effect of a high dislocation density on the formation of GP zones.

At room temperature the bulk diffusion rate of solute in Al is so small that the time frame
for GP zones to form would be orders of magnitude larger than the 3 to 4 days observed.
It is believed that the excess vacancies captured during quenching play a critical role in
the formation of GP zones and that these excess vacancies are responsible for the
increased diffusion rates. If this is accurate, knowing that dislocations act as sinks for
vacancies, it should be possible to introduce a large density of dislocations and inhibit
natural aging.



252
Experimental Procedure

A 2xxx series aluminum alloy that would effectively strain harden and naturally age was
selected. The alloy chosen was 2524 with a composition of 4.5 wt% Cu, 1.5 wt% Mg, 0.5
wt% Mn and was received in the T4 condition. Specimen with dimensions of 32 x 32 x
10.5mm were machined out of the as-received plate, solution treated at 525C for 25 min
and quenched in non-diluted Aqua-Quench 380, which is a polymer solution that
provided a more uniform and slightly slower quench then water. The polymer quenching
solution was used to avoid cracking at the rolling plane during deformation. The samples
were then cross rolled at room temperature and quenched in an ice bath between
reductions to minimize any GP zone formation by adiabatic heating during the rolling
process. Samples were cold rolled to strains of 0.1, 0.4, and 0.7 immediately after
quenching in the polymer solution. The samples deformed to a strain of 0.4 were also
rolled after different periods of natural aging. The microstructural changes were
indirectly monitored by eddy current electrical conductivity and Rockwell B hardness
measurements as a function of time after quenching. Eddy current electrical conductivity
is a commonly used technique in the aluminum industry to monitor phase transformations
and verify processing. This is because solute in solution, second phase particles, and
dislocations all scatter electrons to different degrees, which is reflected in the
conductivity of the alloy.

Results and Discussion

The phenomenon of natural aging has been known to exist and has been exploited
industrially for nearly one hundred years. The hardness and conductivity profiles
associated with the phase transformation are shown in Figures 2a and b. As can be seen
from the natural aging curves there is an initially rapid raise and quickly saturating
increase in hardness and decrease in conductivity associated with the formation of GP
zones. Both the hardness and conductivity plateau; this is know to occur when the excess
vacancies produced by quenching are lost and the diffusion coefficient returns to the very
low value expected at room temperature, effectively stopping the formation of GP zones
[5].

A rolling strain of 0.7 was imposed on a sample immediately after quenching (at time
equal to zero) in order to test the hypothesis that in the presence of a high dislocation
density natural aging will be inhibited. The time interval between quenching and the
completion of cold rolling was approximately 10min (in all experiments). The hardness
and conductivity data at a strain of 0.7 is shown in Figure 2a and b.


253
50
55
60
65
70
75
80
85
90
95
0 20 40 60 80 100 120 140 160 180
Time (h)
H
a
r
d
n
e
s
s

(
R
B
)
CR at 0 h
Naturally Aged

28.0
28.2
28.4
28.6
28.8
29.0
29.2
29.4
29.6
29.8
30.0
0 20 40 60 80 100 120 140 160 180
Time (h)
C
o
n
d
u
c
t
i
v
i
t
y

(
%
I
A
C
S
)
CR at 0 h
Naturally Aged

Figure 2a (left) time plotted against hardness and 2b (right) time plotted against electrical
conductivity for alloy 2524 naturally aged and cold rolled to a strain of 0.7 at time equal
to zero.

Examination of figure 2a reveals a large initial strain hardening response for the
deformed sample, as would be expected from cold rolling of a high solute alloy. Unlike
the naturally aged sample that showed a steady increase in hardness over the first 40h
typical of natural aging, the hardness of the deformed sample stays constant. The absence
of an increase in hardness indicates an absence of natural aging. The deformed sample
also has a much higher final hardness then the naturally aged sample. The initial increase
in hardness in the rolled sample is accompanied by a simultaneous drop in conductivity.
This drop is clearly correlated with the increase in dislocation density. A surprising result
is that there is a recovery in conductivity. This slight recovery may be due to solute
segregation at dislocations, effectively making multiple electron scattering points (solute
and dislocations) into one, producing a rise in conductivity.

At a strain of 0.7 it appears that natural aging has been inhibited. The same experiment
was performed at a strain of 0.1 in order to determine the effect of a lower dislocation
density on the formation of GP zones. The hardness results are shown in Figure 3.

Similar to the sample cold rolled to a strain of 0.7 there is an initial increase in hardness,
although it is much less, as would be expected. Unlike the heaver rolled sample, whose
hardness remained constant as time increased, the sample strained to 0.1 showed an
increase in hardness within the first 24h. This increase in hardness at room temperature is
characteristic of natural aging, but it is only at about 15% of the naturally aged response
in the unreformed alloy. It appears that the dislocation density produced at a strain of 0.1
is not enough to inhibit natural aging.
50
55
60
65
70
75
80
85
90
95
0 20 40 60 80 100 120 140 160 180
Time (h)
H
a
r
d
n
e
s
s

(
R
B
)
CR at 0h
Naturally Aged

Figure 3. Time in hours plotted against Rockwell B hardness and for alloy 2524
naturally aged and immediately cold rolled to a strain of 0.1.

254

An intermediate strain of 0.4 was chosen and the experiment was once again performed.
Samples were not only rolled at time equal to zero, but after different periods of the
natural aging process. Regardless of what point during the aging process the samples
were rolled each showed a very small increase in hardness and no additional hardening
due to continued aging. Similar but smaller conductivity falls were observed after rolling,
with a partial recovery after the initial fall. The specimen rolled after periods of 6.4h and
12.3h of natural aging increase in final hardness, as shown in Figure 4a. This is due to a
combined hardening effect from the formation of GP zones followed by the strain
hardening. There is a change in this trend observed in the sample rolled at 27.8h. The
final hardness is no longer increasing with time; it actually drops at 27.8h, as shown in
Figure 4b.

50
55
60
65
70
75
80
85
90
95
0 20 40 60 80 100 120 140 160 180
Time (h)
H
a
r
d
n
e
s
s

(
R
B
)
Nat ur ally Aged
CR at 0h
CR at 6.4h
CR at 12.3h
50
55
60
65
70
75
80
85
90
95
0 20 40 60 80 100 120 140 160 180
Time (h)
H
a
r
d
n
e
s
s

(
R
B
)
Nat ural l y Aged
CR at 0h
CR at 6.4h
CR at 12.3h
CR at 27.8h
CR at 100.3h

Figure 4a (left) Rockwell B hardness plotted against time in hours for samples rolled
from 0h to 12.3h and 4b (right) from 0h to 160h.

The reason for this change in final hardness as time prior to cold rolling increases is the
reduction in the effectiveness of strain hardening in the presence of GP zones, as shown
in Figure 5a. As time before rolling increased and more GP zones formed, the increment
of hardening due to aging increased but the increment of strain hardening decreased.



255
0
2
4
6
8
10
12
14
16
18
H
a
r
d
e
n
i
n
g

I
n
c
r
e
m
e
n
t

(
R
B
)
0 6 . 4 12. 3 27 . 8 100 . 3 16 0
Time (h)
Nat ural Aging
Cold Rolling

0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
C
o
n
d
u
c
t
i
v
i
t
y

(
%
I
A
C
S
)
0 6.4 12.3 27.8 100 160
Time (h)
EC Drop

Figure 5a (left) Hardening increment due to aging and strain plotted against time and
5b (right) the drop in conductivity plotted against time.

The drop in electrical conductivity upon cold rolling was attributed to the increase in
dislocation density, and it can be seen in Figure 5b that the drop in conductivity decreases
with the formation of GP zones. This reinforces the observation that the alloy more
effectively strain hardened before the solute forms GP zones. The decrease in the
magnitude of conductivity drop can be attributed to more dynamic recovery, resulting in
less of a strain hardening increment. By plotting the data in Figures 5a&b together, the
relationship between the increment of strain hardening and drop in electrical conductivity
is more easily seen, as shown in Figure 6.

0
2
4
6
8
10
12
14
16
18
20
0 20 40 60 80 100 120 140 160 180
Time Prior to Defromation (h)
H
a
r
d
e
n
i
n
g

I
n
c
r
e
m
e
n
t

(
R
B
)
0
0.2
0.4
0.6
0.8
1
1.2
1.4
C
o
n
d
u
c
t
i
v
i
t
y

(
%
I
A
C
S
)
CR
Agi ng
EC Dr op

Figure 6. The increment of hardening due to cold rolling and naturally aging (on the left
axes) and drop in electrical conductivity (on right axes) plotted against time.

The most striking feature in Figure 6 is the similarity in the profiles of the strain
hardening curve and the drop in electrical conductivity curve. This supports the theory
that dynamic recovery is more effectively inhibited by solute in solution then GP zones.
It is also possible that the cold rolling could have partially redissolved the GP zones, by
cutting them to below a critical size. If this had occurred it would be likely to (a) offset
the strain hardening and (b) reduce the fall in conductivity, which could be an alternative
explanation of the current results. To test this possibility the following experiment was
carried out.

After two weeks at room temperature, the specimen cold rolled at time equal to 0 and
after 160h of natural aging was subjected to an additional deformation by cold rolling to a
strain of 0.3 (producing a total strain of 0.7). The sample deformed immediately after
quenching, to inhibit the formation of GP zones, still exhibits a larger increment of strain
hardening and a lager drop in conductivity, as shown in Figure 7. This is evidence that

256
the GP zones have not been significantly put back into solution by either increments of
cold rolling, the initial strain of 0.4 after160h and the later second reduction. It also
indicates that there has been little GP zone formation after extended holding at room
temperature of the sample rolled at time equal to zero, supporting the conclusions
previously drawn from the current results.

0
0.5
1
1.5
2
2.5
3
3.5
4
I
n
c
r
e
m
e
n
t

0 160
D
e
l
t
a


(
%
I
A
S
C
)
D
e
l
t
a

H
R
B
Time (h)

Figure 7. The increment of hardening due to rolling and drop in electrical conductivity
for the samples rolled at time equal to zero and 160h after an additional deformation to a
strain of 0.3.

An unexpected result was that the sample deformed at time equal to zero to a strain of 0.7
in a single deformation had a final hardness of 89 and the sample deformed to a strain of
0.7 in two processes, the second being two weeks later, had a final hardness of 85. It is
unclear why there is a 4 point discrepancy, although it is possibly due to quenching
efficiency, resulting in slightly different amounts of solute in solution at time equal to
zero.

Conclusions

1) Strain hardening is more effective in the presence of solute then GP zones. As
time at room temperature prior to cold rolling increased the increment of
hardening due to GP zones increased and strain hardening became less effective.
The decrease in hardening due to strain was accompanied with decrease in the
drop of electrical conductivity upon deformation. The drop in conductivity is
attributed to dislocation density, so dynamic recovery is less effectively inhibited
in the presence of GP zones.
2) Plastic deformation inhibits the formation of GP zones. It was observed that the
gradual hardening and drop in electrical conductivity associated with the
formation of GP zones, and characteristic of natural aging, did not occur in the
samples deformed immediately after quenching (time equal to zero). This same
effect was also reproduced at different periods of natural aging.
3) There is a minimum amount of deformation that must be imposed to produce the
effect seen in conclusion 2. The specimen that was quenched and cold rolled to a
strain of 0.1 did show a gradual hardening over the first 24h at room temperature.
This means that there is a minimum amount of dislocations that need to be put in
to act as sinks for the excess vacancies formed during quenching.
4) The time interval between the initial deformation process to prevent the formation
of GP zones and subsequent deformation processing is irrelevant to the

257
effectiveness of strain hardening. The specimen deformed immediately after
quenching and after full aging behaved the same during a subsequent deformation
process over two weeks later. There was a small difference in their final hardness
values, but this is possibly due to a loss of solute during quenching.

Acknowledgments

Study was funded by the Air Force Office of Scientific Research (AFOSR).
The alloy was supplied by Alcoa.
The quenching and heat treating material was supplied by Houghton International Inc.
Mechanics of Microstructures group at Drexel University, Department of Materials
Science and Engineering.

References

[1] A. K. Vasudevan and R. D. Doherty "Grain Boundary Ductile Fracture in
Precipitation Hardened Aluminum Alloys" Overview #58. Acta Metall., 35 (1987) 1193-
1219.

[2] R. S. Sanders, S .F. Baumann and H. C. Stumpf, Aluminum Alloys Contemporary
Research and Appluications, Eds A. K. Vasudevan and R. D. Doherty (Academic Press
San Diego, CA 1989 pp66-105.)

[3] Roger D. Doherty and John McBride, Aluminum Alloys for Packaging, Eds. J. G.
Morris, H. D. Merchant, E. J. Westerman and P.L. Morris, (TMS-AIME Warrendale, PA
1993 pp 347-368.)

[4] R. D. Doherty and S. F. Baumann, Aluminum Alloys for Packaging, Eds. J. G. Morris,
H. D. Merchant, E. J. Westerman and P.L. Morris, (TMS-AIME Warrendale, PA 1993 pp
369-391.)

[5] D. Turnbull, H.S. Rosenbaum, and H. N. Treaftis, Act Metall. 8 (1960) 277


258
APPENDIX C. Preliminary Experimental Results


A large portion of the preliminary work conducted in support of the current
project focused on Equal Channel Angular Pressing (ECAP) to produce severely
plastically deformed (SPD) microstructures. It was originally thought that strains
unachievable by cold rolling would be required to produce yield strengths of interest
and/or that if rolling could produce strengths of interest the resulting cross-section of the
work piece would be too thin for testing. ECAP was an attractive process not only
because it allowed for large strains to be imposed, but also because the cross-section of
the work piece is not reduced at increasing strains. The versatility of the ECAP process
allows for many different processing routes to be used by changing the orientation of the
billet between passes. The evolution of the crystallographic texture of the microstructure
during processing was originally investigated because there was believed to be an
opportunity in optimizing the processing route to develop the most beneficial texture.
The ECAP fixture that was designed and built for this study is shown in Figure
C-1 and had a geometry of =90 and =20 which imposed a true strain of 1 per pass.
A mild steel jacket was fabricated to hold the tool steel fixture together which is not
shown in the picture below. All of the processing was performed using a billet with a
cross-section of 6mm x 6mm and utilizing route "A" which holds the orientation of the
billet constant during processing. A billet at three different stages of pressing is shown in
Figure C-2.


259

Figure C-1: Image of ECAP fixture placed in an Instron 58R1127 screw driven load
frame.



Figure C-2: Image of a billet as it is (a) passing through the shear plane, (b) almost
through the shear plane, and (c) pushed past the shear plane by the next billet.
C
10mm
B
A


The first material to be processed using the ECAP fixture was commercial purity
aluminum. The specific alloy processed was P0815. The P is an internal designation used
by Alcoa. The numerals following the P designate that the material has a maximum of
0.08 wt% Si and 0.15 wt% Fe. The P0815 was able to be subjected to 10 passes through
the ECAP fixture producing the yield strengths shown in Table C- 1 and Figure C-3.
The true stress-strain curves for each ECAP strain level are shown in Figure C-4. It is
seen that while P0815 effectively strain hardens, even at strains as large as 10, that the

260
strengths produced are not of much use to a mechanical designer. Note the effect of strain
path change between Figure C-3 and Figure C-4.

Table C- 1: ECAP strain and yield strength data plotted in Figure C-4 in tabular form.
ECAP
Strain
Yield Strength
(MPa)
0 30
2 130
4 150
6 170
10 180


y = 112.55x
0.2127
0
20
40
60
80
100
120
140
160
180
200
0 2 4 6 8 10 12
ECAP Strain
C
o
m
p
r
e
s
s
i
v
e

Y
i
e
l
d

S
t
r
e
s
s

(
M
P
a
)
P0815

Figure C-3: True compressive stress plotted against ECAP strain for commercial purity
P0815.

261
0
50
100
150
200
250
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
True Compressive Strain
T
r
u
e

C
o
m
p
r
e
s
s
i
v
e

S
t
r
e
s
s

(
M
P
a
)
1
ECAP; E=0
ECAP; E=2
ECAP; E=4
ECAP; E=6
ECAP; E=10

Figure C-4: True compressive stress plotted against true compressive strain post ECAP
processing.


The fully annealed P0815 had an average grain size of ~100m and a strong
{112}<111> texture component, as shown in Figure C- 5. The ECAP processing resulted
in a {411}<110> shear texture, as shown in Figure C- 6, and average grain size of ~1m
after 10 passes.



262

Figure C- 5: {001}, {011}, and {111} pole figures for the fully annealed P0815 revealing
a strong {112}<111> single texture component. The ND direction corresponds to long
direction of the billet.




Figure C- 6: {001}, {011}, and {111} pole figures for the P0815 processed to a strain of
10 revealing a {411}<110> shear texture. The ND direction corresponds to long direction
of the billet.


ECAP processing of AA2524 and AA5182 was also performed. As reported in
the main body of this document, quenching directly into water after solution heat treating
resulted in the work piece fracturing during SESH, as shown in Figure C-7. The slower
quench produced by the synthetic quenchent allowed for the processing to be performed
(at this early stage of the work the uphill quenching process had not yet been
investigated). The slower quench (100% polymer solution) also produced a weaker T4
condition. The strengths produced by ECAP processing exceeded those typically

263
available commercially, as shown in Table C-2 and Figure C-8. The values given as
being "typical" are those normally reported (or that can be purchased) for commercial
purity aluminum alloys, 5xxx series, and 2x24. The 2x24 minimum values are for the O-
temper.


H

Figure C-7: Effect of a single pass through the ECAP fixture on AA2524 if quenched into
water or polymer after SHT.



Table C-2: Yield strength values in the annealed or T4 condition for the alloys processed
and those achieved through ECAP processing. The typical range represents the maximum
and minimum values reported by others using traditional processing techniques.
Yield Strength (MPa)
Alloy
Annealed
or T4
Typical range
(min-max)
ECAPed
P0815 30 20 - 170 180
5182 135 40 - 270 400
2524 240 95 - 455 620

Polymer
2
O

264
0
100
200
300
400
500
600
700
800
0 0.1 0.2 0.3 0.4 0.5 0.6
True Compressive Strain
T
r
u
e

C
o
m
p
r
e
s
s
i
v
e

S
t
r
e
s
s

(
M
P
a
)
2524 E=1
5182 E=1
815 E=10

Figure C-8: Compressive stress-strain curves for AA2524, AA5182, and P0815 after
ECAP processing.


As the preliminary work continued, it become apparent that the ECAP process
was going to far exceed the strengths of interest and that the 6mm x 6mm cross-section of
the billets were going to prove impractical for reliably characterizing the fracture
toughness. At this point the decision was made to use cold rolling to process the material
so that smaller strains could be imposed and standard size specimen could be tested. The
initial rolling work was performed on 25mm thick AA2124. The peak aged condition was
achieved by aging at 190C for 12h and the SESH material was cold rolled to a true strain
of 0.18 in the T4 condition. The overall tensile properties of the SESH material were
superior to the aged, as shown in Table C-3 . The toughness was determined using a
single edge bend specimen loaded in three-point bending in accordance with ASTM E

265
399. The testing did not meet the requirements to produce valid toughness measurements
and are values given are not geometry independent.

Table C-3: Tensile and toughness values for AA2124
Processing Peak Aged SESH % Change
Yield Strength (MPa) 453 507 12
% Elongation 12 13 1
UTS (MPa) 551 573 4
Area Under Curve (mm
2
) 62 74 20
Reduction in Area (mm
2
) 22 30 36
Toughness (MPa m
1/2
) 70 71 1


The failure to produce valid toughness measurements at a specimen thickness of
~25mm helped to drive the decision to characterize the toughness using nonlinear
fracture mechanics (NLFM). Another factor that influenced the decision to use NLFM
was that there were serious concerns about the non-uniformity of the strain though the
thickness of the material at the rolling reduction being imposed. The 2124 alloy was not
used for any additional working because large oxide particles were found in the
microstructure.

266
APPENDIX D. Characterization of Texture Produced by SPD During ECAP




Characterization of Texture Produced by Severe Plastic Deformation During
Equal Angular Channel Pressing of Commercial Purity Aluminum

Abstract:
In this study, texture analyses has been performed on commercial purity
aluminum (Al-0.15wt% Fe-0.07wt% Si) processed by equal angular channel pressing
(ECAP) to an absolute strain of 10. The results obtained are compared to previously
published texture data, by other investigators, for similar alloys under similar processing
conditions. It is hypothesized that the discrepancies in the processed textures are related
to the initial (pre-processing) textures of the alloys. There was also a 2-D yield surface
produced for this texture, which depicted very little anisotropy.

Introduction
There has been a great deal of interest in submicron-grained (SMG) metals over
the past decade. Two of the most common ways of achieving a SMG structure are by
either rapid solidification or severe plastic deformation (SPD). Unfortunately, rapid
solidification techniques are both expensive and result in a material with a much higher
porosity then if it had been conventionally cast. Because of this there has been a drive to
develop an effective process to produce SPD metals. The most common ways of
producing these ultrafine-grained structures (UFS) are equal angular channel pressing
(ECAP), cyclic extrusion compression (CEC), and high pressure torsion (HPT).
The majority of literature concerning SPD processing of metals has focused on
ECAP as it is the most likely of the processes to be scaled up for commercial production.
The ECAP process was developed by V.M. Segal in the former Soviet Union in the
1970s. ECAP processing entails pressing a metallic billet (work piece) into an open
channel which intersects a second channel of the same dimensions at an angle that
usually varies between 90 and 135. This results in a shear deformation, which imposes
large strains on the work piece. It is important to note that the cross-sectional area of the
work piece remains the same after processing, unlike traditional methods of deformation
such as cold rolling and simple compression, allowing the billet to be processed multiple
times.
The large portion of literature on ECAP concentrates on either processing
parameters or the resulting grain size and its effect on mechanical properties. There has
been very little reported on texture evolution during ECAP processing. What has been
reported reveals an enormous potential for controlling textures by altering processing
parameters, although many of these investigations have produced conflicting results. In
the present investigation the texture of commercial purity aluminum processed to a strain
of 10 is compared to similar alloys processed under similar conditions with conflicting
texture results. The anisotropy of the texture was also investigated. A 2_d yield surface
was plotted for the crystallographic texture reviling a very low degree of anisotropy.

267
Experimental
A direct-chill cast ingot of commercial purity aluminum (Al-0.15wt% Fe-
0.07wt% Si) was homogenized at 620C for three days followed by a five day heat
treatment at 450C to precipitate out Al
3
Fe particles. The ingot had a twin feather
microstructure with an average grain size on the order of millimeters. A billet with the
dimensions of 6mm x 6mm x 80mm (ND x TD x RD) was machined out of the ingot for
ECAP processing. The billet was pressed through the ECAP fixture twice and
recrystallized at 450C for 1hr to produce an equiaxed microstructure with an average
grain size of 45m. This structure is considered the starting material for the study.
The ECAP fixture used for processing has a geometry of =90 and =20 (with
a second matching relief angle on the opposite side of the channel as well), as shown in
Figure 1, and produces an absolute strain of 1 per pass as given by the strain analyses
performed by Iwahashi et al [1]. All processing was done by route A [2-3], in which there
is no rotation of the billet between passes.


Figure 1. Schematic representation of the ECAP fixture illustrating the 90 ()
intersection of the two channels and the radius of the turn formed by 20 ().
RD
ND
TD
=90
ND
TD
=20
RD

The recrystallized billet was subjected to ten passes thought the ECAP fixture by
route A to impose an absolute strain of 10. Texture analysis was performed on the alloy
using a field emission SEM and electron backscattered diffraction (EBSD) pattern
analyses. The EBSD analysis was performed at a working distance of 22mm, tilt angle of
70, and acceleration voltage of 20KV. Two areas of 200m x 200m were scanned
using a step size of 5m. The fraction of successfully indexed patterns for the
combination of both scans was 99% (3767/3808). The texture analysis was performed by
combining the data from these two scans and using the TexSEM Laboratories (TSL)
software.
The weighted texture data produced by the TSL software was then plugged into
the Los Alamos polycrystal plasticity (Lapp) software to create a yield surface. The
analysis was performed using the Bishop-Hill analysis and the full constant criteria.

268
Results
The recrystallized material had an average grain size of 45m and a texture index
of 32. The ODF for the processed (=10) material is shown in Figure 2. Examination of
the ODF reveals that there is a partial fiber texture concentrated around {411}<110>
present with the strongest intensity at 49.6,74.6,75.0 (Bunge Euler angles). The pole
figures support this, as shown in Figure 3. The {111} pole figure reveals that there is a
{111} plane nearly perpendicular to the ND and TD (although it is off set) and the {011}
pole figure reveals a <011> crystal direction in the RD sample direction. The inverse pole
figures also support the presence of a {411}<110> partial fiber, as seen in Figure 4. The
inverse pole figures are illustrating a strong <011> crystal direction in the RD direction
and high intensity of <111> in the TD and ND sample directions. This is a commonly
reported shear texture know as a B partial fiber, which is {hkl}<110>. The
{411}<110> partial fiber has a volume fraction of 3.4% within a tolerance of 5, 15%
within a tolerance of 10, and 39.7% within a tolerance of 15. The texture index of the
processed material is 1.6.


Figure 2. Orientation Distribution Function (ODF) of commercial purity aluminum
(Al-0.15wt% Fe-0.07wt% Si) ECAP processed to an absolute strain of 10.

1
is horizontal (0 to 360), is vertical (0 to 90),
2
is constant (0 to 90).


Figure 3. {001}, {011}, and {111} pole figures of commercial purity aluminum
(Al-0.15wt% Fe-0.07wt% Si) ECAP processed to an absolute strain of 10.

269


ND TD
RD
Figure 4. ND, TD, and RD inverse pole figures of commercial purity aluminum (Al-
0.15wt% Fe-0.07wt% Si) ECAP processed to an absolute strain of 10.

Texture data on commercial purity aluminum after ECAP processing to strains of
8 or higher have been reported by [4-5]. The alloys investigated in both of these studies
had an aluminum content of 99.5% and were processed through an ECAP fixture with
=90 and >0. The processing routes in these experiments differed from the current
experiment. [4-5] used processing route B
C
(90 rotations between passes); route A (no
rotation) was used in the current study.
The work conducted at the National Sun Yat-Sen University by P. W. Kao et al.
[5] resulted in a random texture after a strain of 8. The inverse pole figures produced by
this group are shown in Figure 5. This group concluded that ECAP processing is an
effective grain refining tool, producing a random textured material at a strain of 8 or
greater.


RD TD
Figure 5. RD and TD inverse pole figures of commercial purity aluminum (99.5%)
ECAP processed to an absolute strain of 8 [5].

The work conducted at Tsinghua University by A. Godfrey et al. [4] found that a
strong texture still existed at strains as high as 10. The pole figure produced by this group
is shown in Figure 6. It should be noted that the paper does not state which pole figure
this is. The work concluded that a strong texture still exists at a strain of 10 and the
texture data supports this.


270

Figure 6. Pole figure of commercial purity aluminum (99.5%) ECAP
processed to an absolute strain of 10 [4].

There were no results reported on the relationship between crystallographic
texture and mechanical properties by either [4 or 5]. The anisotropy analysis, for the
texture produced in the current study, revealed very little difference in yield strength with
respect to loading direction, as shown in Figure 7.

-4
-3
-2
-1
0
1
2
3
4
-4 -3 -2 -1 0
Plastic Region
(
T
D
)

Elastic Region
1 2 3 4
Sigma 11
S
i
g
m
a

2
2
( RD)

Figure 7. 2-D yield surface, with RD plotted against TD, of commercial purity aluminum
(Al-0.15wt% Fe-0.07wt% Si) ECAP processed to an absolute strain of 10.

Discussion
The texture data presented in the current study does not entirely correspond to the
results published by the other investigators. The results published by A. Godfrey et al. [4]
reveal a strong texture at strains as high as 10, while P. W. Kao et al. [5] reports that the
texture becomes randomized at a strain of 8. A possible explanation of the inconsistencies
may be the EBSD analyses. The EBSD data that produced the random texture consisted
of 350 correctly index patterns in an area of only 6.6m x 10m, while the data that
produced the strongly textured pole figure consisted of approximately 6000 correctly
index patterns in a much larger area of 1mm x 1mm. It is possible that the data that
produced the randomly texture inverse pole figures is not representative of the

271
microstructure. However, this does not explain the difference in texture strength between
the present study and that observed by A. Godfrey et al. [4] since both data sets are
believed to be representative of the bulk texture. The current study shows that at a strain
of 10 the material is still textured, although it is not as strong as reported by A. Godfrey
et al. [4]. It is assumed that the pole figure reported by A. Godfrey et al. [4] is the (111)
pole figure. The texture component observed in this figure is similar to that of (111) pole
figure reported in the current study. There is a slight visual difference between the two
because Figure 6 shows only one sample symmetry variant while Figure 3 shows two. So,
the texture elements are similar: They are both B partial fibers, but the intensities are
different.
Possible explanations for the discrepancies in the work conducted by A. Godfrey
et al. [4] and that at Drexel University are the compositions of the alloys and the
processing routes used. The alloy used in the current study was of high purity and
processed by route A, in contrast to a slightly less pure alloy and processing route B
C
.
An aspect of the texture analyses that was not included in any of the published
data was the original texture. It is most likely that the differences observed in the three
analyses (assuming that the random texture observed at National Sun Yat-Sen University
is representative of the bulk texture) are a product of both original texture and processing
route. It was reported by S. Ferrasse et al. [6] that at strains of 4 or higher the influence of
original texture greatly diminishes (but does not disappear completely) and that in
general, texture strength decreases with increasing strain. It was observed that specimen
with weak initial textures, processed by route A and B
C
to strains of 4 or more, resulted
in weak-to-random textures, with route A being more effective at randomization.
However, if the initial texture is strong it was observed that it was possible to produce
medium-to-strong textures at high strains depending on the processing route, this being
especially true for route B
C
. If the material at National Sun Yat-Sen University had a
weak initial texture, and the alloy at Tsinghua University had a strong initial texture, this
may explain the difference in the results. The initial texture of the specimen processed in
the current study was strong, with a texture index of 32.
The anisotropy of the textures produced by [4 and 5] was not reported on by
either investigator. Analysis of a 2-D yield surface was performed on the texture
developed in the current study, revealing very little anisotropy. Careful examination of
the plot reveals that there is a slightly larger yield stress valve if loaded in uniaxial
compression compared to uniaxial tension.

Summary
The ECAP processing of commercial purity aluminum (Al-0.15wt% Fe-0.07wt% Si) by
route A to an absolute strain of 10 produces a partial fiber texture concentrated around
{411}<110>. This is a common shear texture referred to as a B partial fiber. This is in
conflict with texture data that was reported by [5]. The group from Tsinghua University
reported having a similar texture although the intensity was higher. It is believed that the
discrepancies in texture are due to processing route and initial texture of the undeformed
material. In order to perform a proper comparison between texture components in a
deformed specimen it is necessary to have texture information about the material in the
preprocessed condition. The resulting texture is dependent on the initial texture. The

272
anisotropy analyses perform on the texture produced in the current study revealed a
slightly higher yield strength when loaded in uniaxial compression compared to uniaxial
tension although, the degree of anisotropy is relatively small.

References
1. Yoshinori Iwahashi, Jingtao Want, Zenji Horita, Minoru Nemoto and Terence G.
Langdon. Scripta Materialia. 35 (1996) 143-146.
2. V.M. Segal, Mater. Sci. Eng. A 197 (1995) 157-164.
3. S. Ferrasse, V.M. Segal, K.T. Haetwig, R.E. Goforth, Metall. Mater. Tran. A 28A
(1997) 1047.
4. A. Godfrey, W.Q. Coa, Q. Liu, Mater. Sci. Eng. A361 (2003) 9-14.
5. P.L. Sun, P.W. Kao, C.P. Chang. Materials Science and Engineering. A283
(2000) 82-85.
6. S. Ferrasse, V.M Sega, S.R. Kalidindi, F. Alford. Materials Science and
Engineering. A368 (2004) 28-40.


273
APPENDIX E. Representative Optical Micrographs




Representative optical micrographs were taken at magnifications of 100x and
500x. In the micrographs of the TD the RD is horizontal to the page and the ND is
vertical. In micrographs of the RD plane the TD is horizontal to the page and the ND is
vertical. To clarify any confusion regarding nomenclature:
Rolling Direction (RD) Longitudinal Direction (L)
Transverse Direction Long Traverse Direction (LT)
Normal direction Short Traverse Direction (ST)
Refer to Figure 2-39 for a schematic representation of the specimen reference frame.








274

Figure E-1: Micrograph of the TD plane of the T4 material etched with Keller's reagent at
a magnification of 100x.



Figure E-2: Micrograph of the TD plane of the T4 material etched with Keller's reagent at
a magnification of 500x.

275

Figure E-3: Micrograph of the RD plane of the T4 material etched with Keller's reagent at
a magnification of 100x.



Figure E-4: Micrograph of the RD plane of the T4 material etched with Keller's reagent at
a magnification of 500x.

276

Figure E-5: Micrograph of the TD plane of the UA material etched with Keller's reagent
at a magnification of 100x.



Figure E-6: Micrograph of the TD plane of the UA material etched with Keller's reagent
at a magnification of 500x.

277

Figure E-7: Micrograph of the RD plane of the UA material etched with Keller's reagent
at a magnification of 100x.



Figure E-8: Micrograph of the RD plane of the UA material etched with Keller's reagent
at a magnification of 500x.

278

Figure E-9: Micrograph of the TD plane of the T6 material etched with Keller's reagent at
a magnification of 100x.



Figure E-10: Micrograph of the TD plane of the T6 material etched with Keller's reagent
at a magnification of 500x.

279

Figure E-11: Micrograph of the RD plane of the T6 material etched with Keller's reagent
at a magnification of 100x.



Figure E-12: Micrograph of the RD plane of the T6 material etched with Keller's reagent
at a magnification of 500x

280

Figure E-13: Micrograph of the TD plane of the OA material etched with Keller's reagent
at a magnification of 100x.



Figure E-14: Micrograph of the TD plane of the OA material etched with Keller's reagent
at a magnification of 500x.

281

Figure E-15: Micrograph of the RD plane of the OA material etched with Keller's reagent
at a magnification of 100x.



Figure E-16: Micrograph of the RD plane of the OA material etched with Keller's reagent
at a magnification of 500x.

282

Figure E-17: Micrograph of the TD plane of the T4 material cold rolled to a true strain of
0.16 etched with Keller's reagent at a magnification of 100x.



Figure E-18: Micrograph of the TD plane of the T4 material cold rolled to a true strain of
0.16 etched with Keller's reagent at a magnification of 500x.

283

Figure E-19: Micrograph of the RD plane of the T4 material cold rolled to a true strain of
0.16 etched with Keller's reagent at a magnification of 100x.



Figure E-20: Micrograph of the RD plane of the T4 material cold rolled to a true strain of
0.16 etched with Keller's reagent at a magnification of 500x.

284

Figure E-21: Micrograph of the TD plane of the W material cold rolled to a true strain of
0.17 etched with Keller's reagent at a magnification of 100x.



Figure E-22: Micrograph of the TD plane of the W material cold rolled to a true strain of
0.17 etched with Keller's reagent at a magnification of 500x.

285

Figure E-23: Micrograph of the RD plane of the W material cold rolled to a true strain of
0.17 etched with Keller's reagent at a magnification of 100x.



Figure E-24: Micrograph of the RD plane of the W material cold rolled to a true strain of
0.17 etched with Keller's reagent at a magnification of 500x.

286

Figure E-25: Micrograph of the TD plane of the W material cold rolled to a true strain of
0.31 etched with Keller's reagent at a magnification of 100x.



Figure E-26: Micrograph of the TD plane of the W material cold rolled to a true strain of
0.31 etched with Keller's reagent at a magnification of 500x.

287

Figure E-27: Micrograph of the RD plane of the W material cold rolled to a true strain of
0.31 etched with Keller's reagent at a magnification of 100x.



Figure E-28: Micrograph of the RD plane of the W material cold rolled to a true strain of
0.31 etched with Keller's reagent at a magnification of 500x.

288

Figure E-29: Micrograph of the TD plane of the T4 material cold rolled to a true strain of
0.41 etched with Keller's reagent at a magnification of 100x.



Figure E-30: Micrograph of the TD plane of the T4 material cold rolled to a true strain of
0.41 etched with Keller's reagent at a magnification of 500x.

289

Figure E-31: Micrograph of the RD plane of the T4 material cold rolled to a true strain of
0.41 etched with Keller's reagent at a magnification of 100x.



Figure E-32: Micrograph of the RD plane of the T4 material cold rolled to a true strain of
0.41 etched with Keller's reagent at a magnification of 500x.

290
APPENDIX F. Representative Fractographs




Representative fractographs were taken at magnifications of 100x, 500x, 1,000x
and 5,000x. All fractographs are of the RD plane for C(T) specimen tested in the L-T
(loading in the L direction with crack propagation in the LT direction) orientation. In the
fractographs the TD is horizontal to the page and the ND is vertical. To clarify any
confusion regarding nomenclature:
Rolling Direction (RD) Longitudinal Direction (L)
Transverse Direction Long Traverse Direction (LT)
Normal direction Short Traverse Direction (ST)
Refer to Figure 2-39 for a schematic representation of the specimen reference frame and
orientation of nested C(T) specimen.





291

Figure F-1: Fractograph of the RD plane of the material in the T4-temper at
magnification of 100x.



Figure F-2: Fractograph of the RD plane of the material in the T4-temper at
magnification of 500x.


292

Figure F-3: Fractograph of the RD plane of the material in the T4-temper at
magnification of 1,000x.



Figure F-4: Fractograph of the RD plane of the material in the T4-temper at
magnification of 2,000x.


293

Figure F-5: Fractograph of the RD plane of the material in the UA condition at
magnification of 100x.



Figure F-6: Fractograph of the RD plane of the material in the UA condition at
magnification of 500x.


294

Figure F-7: Fractograph of the RD plane of the material in the UA condition at
magnification of 1,000x.



Figure F-8: Fractograph of the RD plane of the material in the UA condition at
magnification of 2,000x.


295

Figure F-9: Fractograph of the RD plane of the material in the T6-temper at
magnification of 100x.



Figure F-10: Fractograph of the RD plane of the material in the T6-temper at
magnification of 500x.


296

Figure F-11: Fractograph of the RD plane of the material in the T6-temper at
magnification of 1,000x.



Figure F-12: Fractograph of the RD plane of the material in the T6-temper at
magnification of 2,000x.


297

Figure F-13: Fractograph of the RD plane of the material in the OA condition at
magnification of 100x.



Figure F-14: Fractograph of the RD plane of the material in the OA condition at
magnification of 500x.


298

Figure F-15: Fractograph of the RD plane of the material in the OA condition at
magnification of 1,000x.



Figure F-16: Fractograph of the RD plane of the material in the OA condition at
magnification of 2,000x.


299

Figure F-17: Fractograph of the RD plane of the T4 material cold rolled to a true strain of
0.16 at magnification of 100x.



Figure F-18: Fractograph of the RD plane of the T4 material cold rolled to a true strain of
0.16 at magnification of 500x.


300

Figure F-19: Fractograph of the RD plane of the T4 material cold rolled to a true strain of
0.16 at magnification of 1,000x.



Figure F- 20: Fractograph of the RD plane of the T4 material cold rolled to a true strain
of 0.16 at magnification of 2,000x.


301

Figure F- 21: Fractograph of the RD plane of the W material cold rolled to a true strain of
0.17 at magnification of 100x.



Figure F-22: Fractograph of the RD plane of the W material cold rolled to a true strain of
0.17 at magnification of 500x.


302

Figure F-23: Fractograph of the RD plane of the W material cold rolled to a true strain of
0.17 at magnification of 1,000x.



Figure F-24: Fractograph of the RD plane of the W material cold rolled to a true strain of
0.17 at magnification of 2,000x.


303

Figure F-25: Fractograph of the RD plane of the W material cold rolled to a true strain of
0.31 at magnification of 100x.



Figure F-26: Fractograph of the RD plane of the W material cold rolled to a true strain of
0.31 at magnification of 500x.


304

Figure F-27: Fractograph of the RD plane of the W material cold rolled to a true strain of
0.31 at magnification of 1,000x.



Figure F-28: Fractograph of the RD plane of the W material cold rolled to a true strain of
0.31 at magnification of 2,000x.


305

Figure F-29: Fractograph of the RD plane of the T4 material cold rolled to a true strain of
0.41 at magnification of 100x.



Figure F-30: Fractograph of the RD plane of the T4 material cold rolled to a true strain of
0.41 at magnification of 500x.


306

Figure F-31: Fractograph of the RD plane of the T4 material cold rolled to a true strain of
0.41 at magnification of 1,000x.



Figure F-32: Fractograph of the RD plane of the T4 material cold rolled to a true strain of
0.41 at magnification of 2,000x.

307
VITA

Name: Christopher James Hovanec

Birth: Sellersville, Pennsylvania, United States of America, July 28th 1980.

Citizenship: United States, holding a Security Clearance

Education:

Drexel University Philadelphia, PA
Doctor of Philosophy in Materials Science and Engineering GPA: 3.9/4.0
April 2011

Drexel University Philadelphia, PA
Bachelor of Science in Materials Engineering GPA: 3.1/4.0
June 2004

Experience:

Materials Engineer (Metallurgist)
Naval Air Warfare Center Aircraft Division, Patuxent River, MD Oct. 2007 - Present

Graduate Research Assistant, Department of Materials Science and Engineering
Drexel University, Philadelphia, PA Aug. 2004 - Oct. 2007

Graduate Teaching Assistant, Department of Materials Science and Engineering
Drexel University, Philadelphia, PA 2006 and 2007

Laboratory Manager and Facilities Representative
Drexel University, Philadelphia, PA Aug. 2005 - Oct. 2007

Advanced Coatings Process Engineer
Pratt & Whitney, East Hartford, CT March 2003Sept. 2003

Failure Analysis/Applied R&D Task Group Member
American Competitiveness Institute, Philadelphia, PA March 2002Sept. 2002

Melt Shop Metallurgist
Bethlehem Lukens Plate, Coatesville, PA April 2001Nov. 2001

Honors, Awards, & Activities:

Graduate Assistance in Areas of National Need (GAANN) Fellow, Koerner Fellow, A. W. and
Dorothy Grosvenor Scholarship recipient, H. H. Harris Foundation Scholarship recipient;

Outstanding Graduate Student Award and Graduate Teaching Award, Manufacturing and
Operations Recognition of Excellence Award;

Student Chair of ASM International Liberty Bell Chapter, Materials Engineering Society of
Professional Organizations (MESPO) Student Treasurer.

308






Page intentionally left black.

Das könnte Ihnen auch gefallen