Вы находитесь на странице: 1из 198

Investigation of Cyclic Deformation and Fatigue of

Polycrystalline Cu under Pure Compression Cyclic


Loading Conditions

By

Tzu-Yin Jean Hsu

A thesis submitted in conformity with the requirements


for the degree of Doctor of Philosophy
Graduate Department of Materials Science and Engineering
University of Toronto

 Copyright by Tzu-Yin Jean Hsu 2015


Investigation of Cyclic Deformation and Fatigue of Polycrystalline Cu
under Pure Compression Cyclic Loading Conditions
Tzu-Yin Jean Hsu

Doctor of Philosophy

Materials Science and Engineering


University of Toronto

2015

Abstract
It is commonly accepted that fatigue crack is initiated under tensile fatigue stresses.

However, practical examples demonstrate that cracks may also initiate under pure

compressive fluctuating loads such as the failures observed in aircraft landing gear frames.

However, the mechanism of such failures is rarely investigated. Furthermore, knowledge on

cyclic deformation response under pure compressive fatigue condition is also very limited or

non-existent. Our recent work already verified that fatigue cracks may nucleate from stress

concentration sites under pure compression fatigue, but whether or not a form of stress

concentration is always needed to initiate a crack under pure compression fatigue remains

uncertain. In this study, compression fatigue tests under different peak stresses were carried

out on smooth bars of fully annealed OFHC Copper. The purpose of these tests is to

investigate not only the cyclic deformation response but also the possibility of crack

nucleation without the stress concentrator. Results showed that overall the cyclic stress-strain

response and microstructural evolution of OFHC Copper under pure compression fatigue

exhibits rather dissimilar behaviour compared to those under symmetrical fatigue. The

specimens hardened rapidly within 10 cycles under pure compression fatigue unlike the

gradual cyclic hardening behaviour in symmetrical fatigue with the same peak stress

amplitude. Compressive cyclic creep behaviour was also observed under the same testing

conditions. Moreover, unlike conventional tension-compression fatigue, only moderate slip

activity was detectable on the surface instead of typical PSB features detected from TEM
II
observations. The surface observations has revealed that surface slip bands did not increase

in number nor did they become more pronounced in height with increasing number of cycles.

In addition, surface roughening by grain boundary extrusion was detected to become more

severe as the cycling progressed. Therefore, it was suggested that the plastic strain

accommodated within the samples was not in any major way related to dislocation activities.

Instead, the cyclic stress-strain response for pure compression fatigue was correlated with

surface morphology evolution. In other words, cyclic creep under pure compression fatigue

was caused mainly by the mechanism of grain boundary extrusion. Such phenomenon was

found to eventually lead to crack nucleation. Furthermore, from symmetrical fatigue testing,

it was noted that depending on the polarity of the loading spectrum in the first half cycle, i.e.

compression or tension, either regular Bauschinger effect or anti-Bauschinger effect was

observed, respectively. Such observations further elucidate the role of compression loading

spectrum versus tension loading in fatigue.

III
Acknowledgements
First and foremost, I would like to dedicate this thesis and express my deepest gratitude

to Professor Zhirui Wang, who was not only my supervisor but also my mentor and my go-to

dad. His guidance and challenge ways have not only brought out the best of myself but also

shaped me to the person who I am today. Words cannot express my appreciation towards all

his kindness, support, and loving care all these years. Thank you, Professor Wang for giving

me this wonderful opportunity to be part of the ―Wang‘s group‖ and for all the wisdom that

you have so selflessly shared as well as treating me as your own. Most importantly, thank

you for being the inspiration to us all.

I am also very grateful to my committee members, Professor Doug Perovic, Professor

Glenn Hibbard, and Professor Tom North for their continuing support in serving on my thesis

committee. A special thank you to Professor Ludvik Kunz for his kind advice and comments

to my thesis.

A sincere thank you to Dr. Charles Kwan, Dr. Jagan Ulaganathan, Mr. Sal Boccia, Dr.

Dan Grozea and Mr. Doug Holmyard for their help on the technical aspects of my project.

Without their help, completion of this project may have been much difficult. I would also

like to express my gratitude to Dr. Da-Wei Yu, Dr. Balaji Devathavenkatesh and Dr. Sanaz

Ketabi, and for their generous offering hand with my experimental torch setup countless

times especially during these late night experiments. Many experiments were achieved

smoothly because of you. To my summer student, Miss Melody Cheng, thank you for your

diligent work. Without you, my graduation date may have been delayed.

To Miss Judy Ue, a warmest thank you from the bottom of my heart for always having

my back and your forever faith in me even when I was in doubt. I wouldn‘t have the courage

to have gone this far without your encouragements and supports. You were truly an

exceptional friend indeed!

To my dear MSE and CHE friends, Dr. Kitty Kumar, Dr. Sanaz Ketabi, , Dr. Mark Li,

IV
Miss Karen Chien, Mr. Khaled Abu Samk, Dr. Balaji DevathavenkateshMr. Ante Lausic,

Miss Megan Hostetter, Ms. Nastaran Abbarin and Dr. Jagan Ulaganathan, thank you for your

friendships, companionships and many sweet memories. You guys were the best surprise

birthday party organizing committee there is. My Ph.D life would not be as eventful as it

were without you. Furthermore, a special appreciation to Professor Hibbard and his group for

making me as their ―honorary member‖.

To MSE department staff, Ms. Maria Fryman, Ms. Jody Prentice, Ms. Fanning Strumas-

Manousos, thank you very much for helping me to keep the administrative aspects of this

project in order all these years.

The funding support from University of Toronto Open Fellowship, US.Steel Scholarship

and Ontario Graduate Scholarship are also greatly appreciated.

Lastly but not the least, I would like to give my sincere gratitude to my parents Mr.

William Hsu and Mrs. Maryann Hsu-Chen and my brother, Mr. Jason Hsu as well as my

Canadian extended family, Mrs. Cathy Bennett, Mrs. Monica Chew, and late Grandpa Don

Matheson for all their constant and unconditional love, moral supports and their unending

believe in me. They have helped me to find my way when I most need it. Most of all, through

the good time and the bad times, they have always there for me! Thank you my dear family, I

love you all always!

V
List of Publications

Based upon the work of this thesis, the following refereed papers have been published

and presented.

Refereed journal papers:

 Tzu-Yin Jean Hsu, Zhirui Wang, ―Cyclic stress-strain response and

microstructure evolution of polycrystalline Cu under pure compressive

cyclic loading condition‖, Material Science and Engineering. A, 615 (2014)

pp.302-312.

 Tzu-Yin Jean Hsu, Zhirui Wang, ―Cyclic deformation response and crack

initiation mechanisms under pure compression fatigue‖, Advance Materials

Research, 891 (2014) pp.470-475

Refereed conference papers:

 Tzu-Yin Jean Hsu, Zhirui Wang, Fatigue crack initiation at notch root under

compressive cyclic loading, Procedia Engineering, 2 (2010) pp.91-100

Manuscript in preparation for refereed journals:

 Tzu-Yin Jean Hsu, Zhirui Wang, ―Investigation of crack nucleation mechanisms

of polycrystalline cu under pure compressive cyclic loading condition‖

 Tzu-Yin Jean Hsu, Zhirui Wang, ―Asymmetrical mechanical response of OFHC

polycrystalline Cu‖

VI
Table of Contents
ABSTRACT ............................................................................................................................ II

ACKNOWLEDGEMENTS ................................................................................................. IV

LIST OF PUBLICATIONS ................................................................................................. VI

TABLE OF CONTENTS ................................................................................................... VII

LIST OF TABLES ................................................................................................................ XI

LIST OF FIGURES ............................................................................................................ XII

LIST OF ACRONYMS AND SYMBOLS ....................................................................... XXI

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

1.1. COMPRESSION FATIGUE – AN OVERVIEW................................................................... 1

1.1.1. Recent Study of Crack Initiation under Pure Compression Fatigue Load ........... 3

1.2. OBJECTIVES ............................................................................................................... 7

1.3. STRUCTURE OF THE THESIS ........................................................................................ 8

2. LITERATURE REVIEW .............................................................................................. 9

2.1. OVERVIEW ................................................................................................................. 9

2.2. TERMINOLOGY OF CYCLIC DEFORMATION RESPONSE ............................................. 11

2.2.1. Cyclic Softening/Hardening Behaviour .............................................................. 11

2.2.2. Cyclic Creep........................................................................................................ 12

2.3. CYCLIC DEFORMATION BEHAVIOUR OF COPPER UNDER SYMMETRICAL FATIGUE ... 13

2.3.1. Monocrystalline Copper ..................................................................................... 14

2.3.2. Polycrystalline Copper ....................................................................................... 19

2.4. CYCLIC DEFORMATION BEHAVIOUR UNDER THE INFLUENCE OF MEAN STRESS ...... 24

2.4.1. Effect of Mean Stress on Cyclic Plasticity .......................................................... 24

2.4.2. Effect of Mean Stress on Fatigue Strength and Fatigue Life .............................. 27

VII
2.5. GENERAL FATIGUE CRACK NUCLEATION MECHANISMS .......................................... 32

2.5.1. Effect of Stress Concentrators ............................................................................ 33

2.5.2. Role of Persistent Slip Bands & Intrusions and Extrusions ............................... 34

2.5.3. Effect of Grain and Twin Boundaries ................................................................. 40

2.6. FATIGUE CRACK BEHAVIOUR UNDER PURE COMPRESSION FATIGUE CONDITIONS .. 43

2.7. ASYMMETRICAL BEHAVIOUR – AN OVERVIEW OF BAUSCHINGER EFFECT ............... 48

2.7.1. Terminology ........................................................................................................ 49

2.7.2. Mechanisms of Bauschinger Effect ..................................................................... 50

2.7.2.1 Dislocation Approach ..................................................................................... 50

2.7.2.2 Composite Model ............................................................................................ 54

3. METHODOLOGY AND EXPERIMENTAL PROCEDURES................................ 57

3.1. EXPERIMENTAL STRATEGY ...................................................................................... 57

3.2. MATERIAL SELECTION ............................................................................................. 57

3.3. SAMPLE DESIGN AND TEST SET-UP .......................................................................... 58

3.4. MICROSTRUCTURE CHARACTERIZATION .................................................................. 61

3.5. MECHANICAL PROPERTIES CHARACTERIZATION ..................................................... 61

3.6. CYCLIC LOADING TESTS .......................................................................................... 62

3.7. SEMI-IN-SITU OBSERVATION AND CHARACTERIZATION OF SURFACE EVOLUTION DUE

TO CYCLIC LOADING ............................................................................................................ 65

3.7.1. Sample Surface Preparation ............................................................................... 65

3.7.2. Micro-hardness Indent Spacing Measurements.................................................. 66

3.7.3. Optical and Scanning Electron Microscopy (including Electron Backscattered

Imaging) .......................................................................................................................... 67

3.7.4. Atomic Force Microscopy Measurement ............................................................ 67

3.8. CHARACTERIZATION OF DISLOCATION EVOLUTION DUE TO CYCLIC LOADING BY

TRANSMISSION ELECTRON MICROSCOPY ............................................................................. 68


VIII
4. RESULTS AND DISCUSSION I: CYCLIC STRESS-STRAIN RESPONSE AND

MICROSTRUCTURE EVOLUTION OF POLYCRYSTALLINE CU UNDER PURE

COMPRESSIVE CYCLIC LOADING CONDITION ...................................................... 69

4.1. OVERVIEW ............................................................................................................... 69

4.2. GENERAL PROPERTIES OF CU ................................................................................... 70

4.2.1. Microstructure Characterization ........................................................................ 70

4.2.2. Mechanical Properties ........................................................................................ 70

4.3. CYCLIC RESPONSE ................................................................................................... 72

4.3.1. Cyclic Stress Strain Response ............................................................................. 73

4.3.2. Cyclic Hardening Curve and Cyclic Stress Strain Curve ................................... 78

4.3.3. Cyclic Creep........................................................................................................ 81

4.4. SURFACE MORPHOLOGY EVOLUTION ...................................................................... 84

4.4.1. Optical and SEM Observations .......................................................................... 84

4.4.2. AFM Observations .............................................................................................. 89

4.5. DISLOCATION EVOLUTION ....................................................................................... 93

4.6. ANALYSIS ON THE CYCLIC STRESS STRAIN RESPONSE............................................. 95

4.7. CORRELATION OF THE MICROSTRUCTURE EVOLUTION AND THE CSS RESPONSE .... 98

4.8. SUMMARY .............................................................................................................. 102

5. RESULTS AND DISCUSSION II: INVESTIGATION OF CRACK

NUCLEATION MECHANISMS OF POLYCRYSTALLINE CU UNDER PURE

COMPRESSIVE CYCLIC LOADING CONDITION .................................................... 104

5.1. OVERVIEW ............................................................................................................. 104

5.2. CYCLIC CREEP ....................................................................................................... 105

5.3. DISLOCATION STRUCTURE ..................................................................................... 107

IX
5.4. GENERAL SURFACE MORPHOLOGY EVOLUTIONS – DETECTION OF CRACK

NUCLEATION...................................................................................................................... 108

5.4.1. AFM Observations ............................................................................................ 109

5.4.2. Crack Detection ................................................................................................ 112

5.5. DISCUSSION ........................................................................................................... 124

5.6. SUMMARY .............................................................................................................. 129

6. RESULTS AND DISCUSSION III: ASYMMETRICAL MECHANICAL

RESPONSE OF OFHC POLYCRYSTALLINE CU....................................................... 131

6.1. OVERVIEW ............................................................................................................. 131

6.2. GENERAL CYCLIC DEFORMATION RESPONSE OF CU UNDER SYMMETRICAL

CONDITIONS ....................................................................................................................... 131

6.3. EVALUATION OF BAUSCHINGER PARAMETERS....................................................... 136

6.4. MECHANICAL PROPERTIES OF CU .......................................................................... 143

6.5. EXPLANATIONS FOR DISCREPANCIES IN BAUSCHINGER EFFECT BETWEEN TWO

LOADING CONDITIONS ....................................................................................................... 144

6.5.1. Analysis of Strain Hardening Rate ................................................................... 144

6.6. TEM OBSERVATIONS ............................................................................................. 148

6.7. SUMMARY .............................................................................................................. 151

7. CONCLUSIONS ......................................................................................................... 152

7.1. SUMMARY OF THE KEY EXPERIMENTAL OBSERVATIONS ....................................... 152

7.2. MAJOR SCIENTIFIC CONTRIBUTIONS ...................................................................... 154

8. RECOMMENDATION FOR FUTURE WORK ..................................................... 156

9. REFERENCES ............................................................................................................ 158

APPENDIX A ...................................................................................................................... 173

X
List of Tables
TABLE 3-1: LIST OF IMPURITIES IN OFHC CU 101 IN PPM [161] .............................................. 57
TABLE 3-2: CHEMICAL COMPOSITION OF WOOD‘S METAL IN WT% [164] ................................ 59
TABLE 3-3: SUMMARY OF FATIGUE TESTING CONDITIONS ........................................................ 63
TABLE 3-4: SUMMARY OF SEMI IN-SITU TRACE CONDITIONS .................................................... 65
TABLE 4-1: TENSILE PROPERTIES............................................................................................. 72
TABLE 4-2: TENSILE VS. COMPRESSION TEST RESULTS ........................................................... 72
TABLE 5-1: LIST OF GRAIN BOUNDARY ANGLE MEASUREMENTS FROM FIGURE 5-14 ........... 123
TABLE 5-2 SUMMARY OF Θ MEASUREMENTS AT GRAIN BOUNDARY EXTRUSION SITES FROM
FIGURE 5-15 .................................................................................................................. 124
TABLE 6-1: SUMMARY OF BAUSCHINGER PARAMETERS FOR DIFFERENT PEAK STRESS
CONDITIONS ................................................................................................................... 136

XI
List of Figures
FIGURE 1-1: LOAD-STROKE CURVE OF LANDING GEAR STRUT [8] .............................................. 2

FIGURE 1-2: CONSIDERATION OF VON MISES STRESS VS. NUMBER OF CYCLES TO CRACK

INITIATION (QUENCHED-TEMPERED CONDITION) [13]......................................................... 5

FIGURE 1-3: ILLUSTRATION OF CRACKING ALONG GRAIN ROTATION IN ANNEALED CONDITION

[13] .................................................................................................................................... 5

FIGURE 1-4: CRACK NUCLEATION IN QUENCH TEMPERED SAMPLE: (V-NOTCH) KT =6.72,

NUMBER OF CYCLES: 17000; PEAK STRESS: 90% σYC [13] ................................................. 6

FIGURE 1-5: SCHEMATIC ILLUSTRATION OF THE EFFECT OF STRESS CONCENTRATION SITE AND

THE RELATION OF CYCLIC DEFORMATION BEHAVIOUR OF A MATERIAL .............................. 7

FIGURE 2-1: ILLUSTRATION OF FATIGUE ENDURANCE LIMIT ...................................................... 9

FIGURE 2-2: ILLUSTRATION OF CYCLIC SOFTENING BEHAVIOUR UNDER (A) STRESS CONTROLLED,

AND (B) TOTAL STRAIN CONTROLLED COMPRESSION-COMPRESSION TESTS. THE REVERSED

TREND IS OBSERVED FOR THE CYCLIC HARDENING BEHAVIOUR........................................ 12

FIGURE 2-3: SCHEMATIC OF CYCLIC CREEP BEHAVIOUR UNDER (A) STRESS CONTROLLED, AND

(B) TOTAL STRAIN CONTROLLED COMPRESSION-COMPRESSION TESTS. ............................. 13

FIGURE 2-4: CYCLIC STRESS STRAIN CURVE [14] ..................................................................... 15

FIGURE 2-5: THREE-DIMENSIONAL TEM MICROGRAPH OF THE DISLOCATION ARRANGEMENT IN

MONOCRYSTALLINE CU IN DIFFERENT CYCLIC SATURATION REGION (A) REGION A: γAP =

2.6 × 10-5 (B) REGION B: γAP = 1.5 × 10-3 (C) REGION C: γAP = 1.45 × 10-2 ...................... 16

FIGURE 2-6: DISLOCATION ARRANGEMENTS IN PERSISTENT SLIP BANDS IN FCC METALS [34] 17

FIGURE 2-7: TEM OF A SECTION PARALLEL TO THE PRIMARY SLIP PLANE OF SINGLE CRYSTAL

CU FATIGUE TO SATURATION AT γAP = 5 × 10-3 [35] .......................................................... 17

FIGURE 2-8: COMPARISON OF THE CYCLIC STRESS-STRAIN CURVES. THE SHEAR STRESS IS

TRANSPOSED BY THE TAYLOR FACTOR [19]. .................................................................... 22

XII
FIGURE 2-9: CYCLIC STRESS-STRAIN CURVE OF ANNEALED COPPER SHOWN IN

SEMILOGARITHMIC PLOT ADOPTED FROM [16] ................................................................. 22

FIGURE 2-10: TYPICAL LADDER-LIKE PSBS IN BULK OF POLYCRYSTALLINE: (A) STRADDLING A

TWIN BOUNDARY; (B) CONTAINED ENTIRELY WITHIN A NARROW TWIN; SPECIMEN RAMP-

LOADED TO 98MPA, AND THEN STEP-TESTED IN STRAIN CONTROL TO THE MIDPOINT OF

THE PLATEAU AT 98MPA [19] .......................................................................................... 23

FIGURE 2-11: DISLOCATION CELL STRUCTURE OBSERVED IN MONOCRYSTALLINE CU UNDER

PULSATING TENSION FATIGUE [75] ................................................................................... 26

FIGURE 2-12: COARSE SLIP BANDS ON SURFACE OF MONOCRYSTALLINE CU UNDER PULSATING

TENSION FATIGUE [75] ..................................................................................................... 27

FIGURE 2-13: STRESS REQUIREMENTS FOR THE FORMATION OF PSBS IN SINGLE-SLIP-ORIENTED

COPPER SINGLE CRYSTALS (SCHEMATIC) .......................................................................... 27

FIGURE 2-14: (A) TYPICAL STRESS AMPLITUDE-LIFE PLOTS FOR DIFFERENT MEAN STRESS

VALUES ADOPTED FROM HERTZBERG [78] (B) GERBER, GOODMAN, AND SODERBERG

DIAGRAMS (R-M DIAGRAM) SHOWING COMBINED EFFECT OF ALTERNATING STRESS

AMPLITUDE AND MEAN STRESS ON FATIGUE ENDURANCE [53-55] ................................... 28

FIGURE 2-15: R-M DIAGRAM SHOWING THE EFFECT OF TENSION AND COMPRESSION MEAN

STRESS ADOPTED FROM FORREST [86] ............................................................................. 30

FIGURE 2-16: EFFECT OF MEAN-STRESS ON FRACTURE MECHANISM IN STRESS-CONTROLLED

TESTS [88] ........................................................................................................................ 31

FIGURE 2-17: SOME PSB PROTRUSIONS WITH SUPERIMPOSED EXTRUSIONS AND INTRUSIONS IN

A COPPER SINGLE CRYSTAL. A CRYSTAL TESTED AT A PLASTIC STRAIN AMPLITUDE OF 2 X

10-3 FOR 120,000CYCLES [102]. ...................................................................................... 35

FIGURE 2-18: MODEL OF CARD SLIP IN FATIGUE SLIP BAND [24]. ............................................. 36

FIGURE 2-19 : LADDER LIKE STRUCTURES AND STAGE I CRACKS WITHIN THEM [106].............. 37

XIII
FIGURE 2-20: SCHEMATIC REPRESENTATION OF THE MECHANISM OF EXTRUSIONS/PROTRUSIONS

AT THE PSB ACCORDING TO ESSMANN ET AL. [109] ........................................................ 37

FIGURE 2-21: NEUMANN‘S MODEL OF CRACK NUCLEATION. IN PART (C) A REPRESENTS A

CRACK NUCLEUS [111] ..................................................................................................... 39

FIGURE 2-22: A) A SECTION THROUGH A PSB CONTAINING INTRUSIONS AT A AND B. NOTE THE

CRACK HAS STARTED AT B [112] B) FATIGUE CRACK INITIATION (DENOTED BY THE

ARROW) AT PSB-MATRIX INTERFACE IN A CU CRYSTAL FATIGUE FOR 60,000 CYCLES AT

γPL = 0.002 AT 20°C [113] ................................................................................................ 40

FIGURE 2-23: A) NUCLEATION OF FLAWS ALONG GRAIN BOUNDARY [117] B) WHITE LIGHT

INTERFEROGRAMS SHOWING SLIP STEPS FORMATION AT GRAIN BOUNDARY IN FATIGUE CU.

THE DARK DIAGONAL LINES PARALLEL TO THE ARROW ARE FIDUCIAL MARKERS WHOSE

SEPARATION IS 100 μM [116]............................................................................................ 41

FIGURE 2-24: RELATIONSHIP OF STRESS AXIS AND THE TWIN PLANE (A) STRESS AXIS NORMAL

TO A TWIN BOUNDARY (B) STRESS AXIS INCLINE TO THE TWIN BOUNDARY. THIS INDICATES

THAT THE STRESS AXIS IS CRYSTALLOGRAPHICALLY DIFFERENT IN EACH CRYSTAL [122] 42

FIGURE 2-25: STACK OF TWIN BOUNDARIES. THE WHITE ARROWS INDICATE TWIN BOUNDARIES

[123] ................................................................................................................................ 43

FIGURE 2-26: DEMONSTRATION OF COMPRESSION FATIGUE CRACK IN CT SPECIMEN .............. 45

FIGURE 2-27: ILLUSTRATION OF TERMINOLOGY OF BAUSCHINGER EFFECTS ADOPTED FROM

ABEL [137] ...................................................................................................................... 50

FIGURE 2-28: SCHEMATIC DIAGRAMS SHOWING (A) THE DISLOCATION PILE-UPS AGAINST A

GRAIN BOUNDARY AND (B) THE INTERACTION BETWEEN A MOBILE DISLOCATION AND THE

SECOND PHASE PARTICLE AT DIFFERENT LOADING STAGE [140, 141] ............................... 51

FIGURE 2-29: SCHEMATIC OF STRESS STRAIN RESPONSE OF THE COMPOSITE MODEL [160] ...... 56

FIGURE 3-1: DUMBBELL SHAPED SPECIMEN [162].................................................................... 59

FIGURE 3-2: SCHEMATIC OF SAMPLE DESIGN 1 ........................................................................ 59

XIV
FIGURE 3-3: SELF-ALIGNMENT GRIP FOR MECHANICAL TESTING [163] .................................... 59

FIGURE 3-4: SCHEMATIC OF SAMPLE DESIGN 2 ........................................................................ 60

FIGURE 3-5: MTS ELECTRO-SERVO HYDRAULIC SYSTEM ......................................................... 61

FIGURE 3-6. ILLUSTRATION OF MICROHARDNESS INDENT SPACING MEASUREMENTS ............... 67

FIGURE 3-7. TEM SPECIMENS PREPARED FROM (A) DESIGN 1 & (B) DESIGN2 ......................... 68

FIGURE 4-1: MICROSTRUCTURE OF FULLY ANNEALED OFHC CU 101 ..................................... 70

FIGURE 4-2: TENSILE AND COMPRESSION PROFILES FOR FULLY ANNEALED POLYCRYSTALLINE

OFHC CU 101 ................................................................................................................. 71

FIGURE 4-3. SUMMARY OF HYSTERESIS LOOP DEVELOPMENT IN COMPRESSION FATIGUE FOR

PEAK STRESS CONDITIONS OF (A) 100CF, (B) 120CF, (C) 140CF AND (D) 160CF ............ 74

FIGURE 4-4. ILLUSTRATION OF EVOLUTION FOR 160CF AT (A) CYCLE 1 (B) CYCLE 2 (C)

CYCLE 5 (D) CYCLE 10 ...................................................................................................... 75

FIGURE 4-5. COMPARISON OF HYSTERESIS LOOP BETWEEN SYMMETRICAL FATIGUE AND

COMPRESSION FATIGUE FOR PEAK STRESS CONDITIONS OF 120 % AT DIFFERENT CYCLES.

(A) CYCLE 1 (B) CYCLE 5 (C) CYCLE 10 AND (D) CYCLE 10,000 ........................................ 76

FIGURE 4-6. ILLUSTRATION OF EVOLUTION FOR 160CF AT (A) CYCLE 1 (B) CYCLE 2 (C)

CYCLE 5 (D) CYCLE 10 AND (E) – (H) THE RESPECTIVE MAGNIFIED SELECTED AREA ......... 77

FIGURE 4-7. SUMMARY OF CYCLIC HARDENING RESPONSE FOR (A) SELECTED PURE

COMPRESSION FATIGUE CONDITIONS, AND (B) SYMMETRICAL FATIGUE CONDITIONS ....... 80

FIGURE 4-8. CSSC OF POLYCRYSTALLINE CU UNDER PURE COMPRESSION FATIGUE ................ 81

FIGURE 4-9. CYCLIC CREEP CURVES OF SELECTED CONDITIONS PLOTTED AGAINST TIME ......... 82

FIGURE 4-10. TOTAL CYCLIC CREEP STRAIN AGAINST PEAK STRESS CONDITIONS WITH AND

WITHOUT THE 1ST CYCLE ................................................................................................. 83

FIGURE 4-11. εPL IN THE 1ST CYCLE .......................................................................................... 83

XV
FIGURE 4-12. OPTICAL IMAGES REVEALING THE SURFACE MICROSTRUCTURES EVOLUTIONS OF

100CF SAMPLE AT DIFFERENT NUMBER OF CYCLES AT LOW MAGNIFICATION. (A) 0 CYCLE

(B) 1 CYCLE (C) 100,000 CYCLES...................................................................................... 85

FIGURE 4-13. OPTICAL IMAGES REVEALING THE SURFACE MICROSTRUCTURES EVOLUTIONS OF

120CF SAMPLE AT DIFFERENT NUMBER OF CYCLES AT LOW MAGNIFICATION. (A) 0 CYCLE

(B) 100 CYCLES (C) 100,000 CYCLES ................................................................................ 85

FIGURE 4-14. OPTICAL IMAGES REVEALING THE SURFACE MICROSTRUCTURES EVOLUTIONS OF

160CF SAMPLE AT DIFFERENT NUMBER OF CYCLES AT LOW MAGNIFICATION. (A) 0 CYCLE

(B) 100 CYCLES (C) 100,000 CYCLES ................................................................................ 86

FIGURE 4-15. OPTICAL IMAGES REVEALING THE SURFACE MICROSTRUCTURES EVOLUTIONS OF

250CF SAMPLE AT DIFFERENT NUMBER OF CYCLES AT LOW MAGNIFICATION. (A) 0 CYCLE

(B) 1 CYCLE (C) 10,000 CYCLES........................................................................................ 86

FIGURE 4-16. SEM IMAGES REVEALING THE SURFACE MICROSTRUCTURES EVOLUTIONS OF

160CF SAMPLE AT DIFFERENT NUMBER OF CYCLES AT HIGH MAGNIFICATION FOR SITE NO.

1. (A) 100 CYCLES (B) 15,000 CYCLES (C) 100,000 CYCLES (NOTE: THE DEVELOPMENT OF

THE GRAIN BOUNDARY EXTRUSION AND THE LACK OF MAJOR DEVELOPMENT OF SLIP

BANDS) ............................................................................................................................ 87

FIGURE 4-17. SEM IMAGES REVEALING THE SURFACE MICROSTRUCTURES EVOLUTIONS OF

160CF SAMPLE AT DIFFERENT NUMBER OF CYCLES AT HIGH MAGNIFICATION FOR SITE NO.

2. (A) 5000 CYCLES (B) 15,000 CYCLES (C) 100,000 CYCLES (NOTE: THE GRAIN BOUNDARY

EXTRUSION) ..................................................................................................................... 88

FIGURE 4-18. SEM IMAGES REVEALING THE SURFACE MORPHOLOGY OF 250CF SAMPLE AT

150,000 CYCLES (NOTE: THE RIGHT HAND IS THE FRAMED REGION IN (A) AT A HIGHER

MAGNIFICATION) .............................................................................................................. 88

FIGURE 4-19. ILLUSTRATION OF GRAIN BOUNDARY EXTRUSION .............................................. 90

XVI
FIGURE 4-20. AFM PROFILE RESULTS AND CORRESPONDING IMAGES AT A GRAIN BOUNDARY:

(A)-(E): THE PROFILE MEASUREMENTS AT DIFFERENT NUMBERS OF CYCLES AFTER CYCLE

NUMBER OF 1, 100, 1000, 10,000 AND 100,000 CYCLES, RESPECTIVELY; (F-J) THE

CORRESPONDING IMAGES. ................................................................................................ 91

FIGURE 4-21. (A) MEASUREMENTS OF GRAIN BOUNDARY OFFSET HEIGHT VS. CYCLE NO. , AT

AFM OBSERVATION SPOTS 1 AND 2, AND (B) SLIP BAND OFFSET HEIGHT MEASUREMENTS

VS. CYCLE NO., AT SPOT 4 ................................................................................................ 92

FIGURE 4-22. DISLOCATION STRUCTURE (A) BEFORE FATIGUE TESTING, (B) AFTER FATIGUE

TESTED FOR 1 CYCLE AT 160CF, (C) AFTER FATIGUE TESTED AT 160CF TO 10,000 CYCLES,

(D) AFTER FATIGUE TESTED AT 100CF TO 10,000 CYCLES ............................................... 94

FIGURE 4-23. CYCLIC HARDENING/SOFTENING CURVES OF SAMPLES TESTED IN PURE TENSION

FATIGUE [75].................................................................................................................... 96

FIGURE 4-24. CYCLIC CREEP CURVES OF SAMPLES TESTED IN PURE TENSION FATIGUE [75] ..... 98

FIGURE 4-25. TYPICAL PERSISTENT SLIP BANDS MORPHOLOGY ON THE SURFACE [19] ........... 100

FIGURE 5-1: CYCLIC CREEP CURVES VS. NUMBER OF CYCLES ................................................. 106

FIGURE 5-2:DISLOCATION MORPHOLOGY OF A) 250CF AFTER 1,000,000 CYCLES B) 332CF

AFTER 100,000 CYCLES .................................................................................................. 108

FIGURE 5-3 AFM PROFILE RESULTS AND CORRESPONDING IMAGES AT A GRAIN BOUNDARY: (A)-

(D): THE PROFILE MEASUREMENTS AT DIFFERENT NUMBERS OF CYCLES AFTER 1, 100,

5000, AND 100,000 CYCLES, RESPECTIVELY; (E-H) THE CORRESPONDING IMAGES OF THE

GRAIN BOUNDARY. ......................................................................................................... 110

FIGURE 5-4: AFM RESULTS OF 100CF SHOWING (A) GRAIN BOUNDARY OFFSET HEIGHT

MEASUREMENTS VS. CYCLE NUMBER AT 2 DIFFERENT LOCATIONS AND (B) SLIP BAND

OFFSET HEIGHT MEASUREMENTS VS. CYCLE NUMBERS AT 2 DIFFERENT LOCATIONS ...... 111

FIGURE 5-5: POSSIBLE CRACK SITE ON 120CF AFTER 15,000 CYCLES: (A) LOW MAGNIFICATION

800X (B) HIGHER MAGNIFICATION 3000X....................................................................... 112

XVII
FIGURE 5-6: TWO POSSIBLE CRACK SITES ON 160CF AFTER 100,000 CYCLES. (A) SITE 1 AT A

LOW MAGNIFICATION 2000X, (B) SITE 1 AT A HIGHER MAGNIFICATION 5000X, (C) SITE 2

AT A LOW MAGNIFICATION, 2000XAND (D) SITE 2 AT A HIGHER MAGNIFICATION 3000X 113

FIGURE 5-7: SITE 1 OF 120CF CONDITION IN FIGURE 5-5 AFTER 200,000 CYCLES (A) AT 8000X

(B) AT 20,000X AT THE WHITE CIRCULAR SITE IN (A)...................................................... 114

FIGURE 5-8: OPTICAL TRACE OF A SPECIFIC SITE WHERE A SURFACE OFFSET IS DEVELOPED

ALONG GRAIN BOUNDARIES AT DIFFERENT NUMBER OF CYCLES FOR 250CF: (A) 0 CYCLE

(B) 1 CYCLE (C) 10,000 CYCLE (D) 150,000 CYCLES WITH RECORDED MEAN STRAIN

MARKED AS εPL ON THE TOP LEFT. ................................................................................... 115

FIGURE 5-9 SEM TRACE OF THE SITE CORRESPOND TO CIRCLED REGION IN FIGURE 5-8 AT

DIFFERENT NUMBER OF CYCLES: (A) 10,000 CYCLE @ 5000X (B) 50,000 CYCLE @ 5000X

(C) 150,000 CYCLE @ 5000X (D) 10,000 CYCLES @ 20,000X, (E) 50,000 CYCLES @

20,000X (F) 150, 000 CYCLES @ 20,000X WITH RECORDED MEAN STRAIN MARKED AS ΕPL

ON THE TOP LEFT. ........................................................................................................... 116

FIGURE 5-10 CRACK NUCLEATION SITE ON 250CFAND ITS DEVELOPMENT FROM (A) 50,000

CYCLES @ 6000X, (B) 150,000 CYCLES @ 6000X, AND (C) 1,000,000 CYCLES @ 6000X

AND (D) AT THE LARGER MAGNIFICATION AT 1,000,000 CYCLES @ 10,000X. WITH

RECORDED MEAN STRAIN SHOWN AS εPL. ........................................................................ 117

FIGURE 5-11 TWIN BOUNDARY PROTRUSION SITE IN CONDITION OF 332CF AFTER

100,000CYCLES: (A) OPTICAL AND (B) SEM MICROGRAPH OF THE CIRCLED REGION IN (A)

AT HIGHER MAGNIFICATION @ 10,000X ......................................................................... 119

FIGURE 5-12 (A) SEM IMAGE AND (B) THE EBSD ANALYSIS OF TWIN BOUNDARY EXTRUSION

SITES FOR 332CF CONDITION AFTER 100,000 CYCLES. NOTE: THIS IS THE SAME SITE AS

SHOWN CIRCLED IN FIGURE 5-11. IN PARALLEL, (C) AND (D) SHOW THE SEM AND EBSD

ANALYSIS FOR THE 250CF CONDITION AFTER 1,000,000 CYCLES. ................................. 120

XVIII
FIGURE 5-13. TWO EBSD MAPS TAKEN AT A CRACK SITE FOR 250CF CONDITION AFTER

1,000,000 CYCLES (A) LOCAL MISORIENTATION MAP AND (B) SCHMID FACTOR ALONG

WITH THE GRAIN BOUNDARY CHARACTER MAP .............................................................. 121

FIGURE 5-14: ILLUSTRATION OF GRAIN BOUNDARY ANGLE MEASUREMENTS AT GRAIN

BOUNDARY PROTRUSION SITES FOR CONDITION 120CF. A TOTAL OF 9 PROTRUDED GRAIN

BOUNDARIES ANGLES, , AGAINST THE LOADING AXIS WERE MEASURED. ...................... 123

FIGURE 5-15 DISTRIBUTION OF PROTRUDED GRAIN BOUNDARY ANGLE WITH THE LOADING AXIS.

....................................................................................................................................... 123

FIGURE 5-16 ILLUSTRATION OF DIFFERENT CASES OF GRAIN BOUNDARY ORIENTATION IN

RELATION TO THE SPECIMEN SURFACE AND TO THE LOADING AXIS WHEN (A) A GRAIN

BOUNDARY HAS ITS SURFACE TRACE PERPENDICULAR TO THE LOADING AXIS BUT IT MAY

TILT AT DIFFERENT ANGLES IN THE DEPTH DIRECTION, AND (B) GRAIN BOUNDARIES THAT

O
HAVE DIFFERENT SURFACE TRACE ANGLE TO THE LOADING AXIS. P1 IS AT 70 , WHERE P2
O
IS AT 90 BUT BOTH AT THE SAME TILT ANGLE OF 45° IN THE DEPTH DIRECTION. ........... 127

FIGURE 6-1: SUMMARY OF HYSTERESIS LOOPS FOR PEAK STRESS CONDITION OF 100% σY FOR

(A) TENSION-START (100TCF) AND (B) COMPRESSION-START (100CTF) ....................... 133

FIGURE 6-2: SUMMARY OF HYSTERESIS LOOPS FOR PEAK STRESS CONDITION OF 120%σY FOR

(A) TENSION-START (120TCF) AND (B) COMPRESSION-START (120CTF) ....................... 134

FIGURE 6-3: MEAN STRAIN ANALYSIS OF SELECTED CONDITIONS PLOTTED AGAINST CYCLE

NUMBER ......................................................................................................................... 135

FIGURE 6-4: MEAN STRAIN ANALYSIS OF SELECTED CONDITIONS PLOTTED AGAINST CYCLE

NUMBER FROM WEISS ET AL. [22] .................................................................................. 135

FIGURE 6-5: BAUSCHINGER EFFECT EVALUATION FOR PEAK STRESS CONDITION OF 90% σY FOR

(A) TENSION-START (90TCF) AND (B) COMPRESSION-START (90CTF) ........................... 137

FIGURE 6-6: BAUSCHINGER EFFECT EVALUATION FOR PEAK STRESS CONDITION OF 100% σY FOR

(A) TENSION-START (100TCF) AND (B) COMPRESSION-START (100CTF) ....................... 138

XIX
FIGURE 6-7: BAUSCHINGER EFFECT EVALUATION FOR PEAK STRESS CONDITION OF 120% σY FOR

(A) TENSION-START (120TCF) AND (B) COMPRESSION-START (120CTF) ....................... 139

FIGURE 6-8: YIELD STRENGTHS COMPARISONS ...................................................................... 140

FIGURE 6-9: BAUSCHINGER EFFECT EVALUATION FOR PEAK STRESS CONDITION OF 120%σY FOR

(A) TENSION-START (120TCF) AND (B) COMPRESSION-START (120CTF) ....................... 141

FIGURE 6-10: BAUSCHINGER EFFECT EVALUATION FOR PEAK STRESS CONDITION OF 120%σY

FOR (A) TENSION-START (120TCF) AND (B) COMPRESSION-START (120CTF) ................ 142

FIGURE 6-11: COMPARISON OF ENGINEERING STRESS STRAIN CURVES BETWEEN TENSILE AND

COMPRESSION TESTS ...................................................................................................... 144

FIGURE 6-12: STRAIN HARDENING RATE AS A FUNCTION OF TRUE STRAIN FOR CONDITIONS (A)

100TCF (B) 100CTF...................................................................................................... 146

FIGURE 6-13: COMPARISON OF STRAIN HARDENING RATE FOR THE FORWARD STROKE BETWEEN

100TCF AND 100CTF ................................................................................................... 147

FIGURE 6-14. DISLOCATION STRUCTURE OF CONDITIONS (A) 90TCF (B) 90CTF (C)120TCF

AND (D) 120CTF ............................................................................................................ 150

Figures in Appendix

FIGURE A 1: STRAIN HARDENING RATE AS A FUNCTION OF TRUE STRAIN FOR CONDITIONS (A)

90TCF (B) 90CTF.......................................................................................................... 173

FIGURE A 2: STRAIN HARDENING RATE AS A FUNCTION OF TRUE STRAIN FOR CONDITIONS (A)

120TCF (B) 120CTF...................................................................................................... 174

FIGURE A 3: COMPARISON OF STRAIN HARDENING RATE FOR THE FORWARD STROKE BETWEEN

TCF AND CTF FOR CONDITIONS OF (A) 90%σY (B) 120%σY ........................................... 175

XX
List of Acronyms and Symbols
△K stress intensity factor range
R cyclic stress range

compression yield strength


or (Y.S) yield strength
OFHC oxygen free high conductivity
S-N stress and number of cycles
N number of cycles
εpl cyclic plastic strain
εmean cyclic mean strain

or σa stress amplitude

or strain amplitude
, total strain range
CSSC cyclic stress strain curve
γap or γpl plastic shear strain amplitude
γap,PSB local plastic shear strain amplitude acting in the PSBs
γap,M local plastic shear strain amplitude acting in the matrix
τs or shear stress
PSB persistent slip band
f PSB volume fraction of persistent slip band
M orientation factor for Schmid, Sachs and Taylor models
predicted stress range from Schmid, Sachs and Taylor models

predicted strain range from Schmid, Sachs and Taylor models


maximum stress
minimum stress
or mean stress
σfat fatigue strength
σts tensile strength

XXI
, fatigue strength coefficient
fatigue exponent in Basquin‘s fatigue relation
fatigue life
fatigue ductility coefficient
c fatigue ductility coefficient in Coffin-Manson equation
SWT Smith, Watson and Topper
material‘s constant in Walker‘s equation
Kf fatigue notch factor
Kt stress concentration factor in relation to the applied load
Kmax maximum stress intensity factor
σend,unnotched fatigue endurance limit of unnotched sample
σend, notched fatigue endurance limit of notched sample
q notch-sensitivity-factor
r plastic zone induced ahead of notch tip
CT compact tension
εR reverse strain
or σpeak maximum peak stress
β Bauschinger strain
∆σp permanent softening
σR reverse yield stress or reverse flow stress
forward flow stress
forest hardening effect
back stress expressed in normal stress
εp prestrain
applied stress expressed in shear stress
frictional stress on the gliding plane
back stress
OM optical microscopy
SEM scanning electron microscopy

XXII
AFM atomic force microscopy
TEM transmission electron microscopy
MTS Material Testing Systems
E elastic modulus
UTS ultimate tensile strength
EL% percent elongation
CF compression fatigue
TCF symmetrical fatigue conditions with tension-start
CTF symmetrical fatigue conditions with compression-start
EBSD electron backscatter diffraction
FA fully annealed

unidirectional plastic strain

creep strain in the first cycle


GB grain boundary
θ orientation angle
σy,0.025% forward yield stress at ε = 0.025%
σR, 0.025% reverse yield stress at ε = 0.025%
εPl,tension plastic strain measured in tension stroke
εPl,compression plastic strain measured in compression stroke
εeng engineering strain
εtrue true strain

XXIII
1. Introduction

1.1. Compression Fatigue – an Overview

Fatigue failure by definition is failure that occurs under dynamic loading conditions,

whereby the fluctuating stresses may be much lower than the material‘s strength. It is often

found that failures in materials are fatigue-related [1-3]. From dramatic examples such as the

fuselage failure of China Airlines aircraft in 2002, to a common everyday life example, such

as cracking of the suspension coil in an automobile, they all showed fatigue was the

dominating cause of failure [2, 4]. In fact, it is stated that fatigue failure make up to 90% of

all service failures due to mechanical causes. As fatigue failure is often both catastrophic and

insidious, ever since fatigue phenomenon was discovered in 1830 [5], scientists, researchers

and designers have made considerable efforts in understanding fatigue in materials in order

to take preventive measures for these failures. These include understanding the fundamentals

of cyclic deformation behaviour and uncovering the underlying mechanisms; in addition,

how they eventually lead to crack nucleation and to the final failures.

To date, much of these explorations have been carried out under the conditions of

tension-compression fatigue or pure cyclic tension fatigue. The consideration of pure

compression fatigue, for instance, has always been overlooked. The reasoning was given on

the premises that pure compressive fatigue or the compression portion of the load spectrum

does not promote fatigue crack initiation or propagation. In other words, it is generally

assumed that there is no crack nucleation or cracks remain closed during such conditions [6].

Even according to ASTM E647, the compression part of loading is disregarded in calculating

the stress intensity factor range, △K [7]. In practical applications, however, it is seen that

fatigue cracks may also initiate and grow under pure compressive fluctuating loads. To list a

few examples, such as in the case of landing gear shock strut, an engineering structure that is

designed to absorb the ground-exerted loads during taxi, take-off and landing. These service

modes introduce loading spectrums that are purely compressive as shown in Figure 1-1[8]. It

1
is observed that from time to time landing gear frame fails under such loading conditions[9].

Also, in either a gear of a rolling mill or a main beam of an airplane [10, 11], fatigue cracks

are observed to form from regions of stress concentration in engineering components. It is

also seen that in the very thin specimens of plane-stress status, such cyclic compressive

loading conditions may also lead to catastrophic failures [12]. Consequently, an

understanding of compressive stress fatigue becomes a necessity for efficient design for a

wide range of engineering situations. Specifically, investigations are required to uncover the

mechanisms of cyclic deformation and the fatigue crack initiation that follows under cyclic

compression condition.

Figure 1-1: Load-stroke curve of landing gear strut [8]

This has undoubtly led to the start of this project, the investigation of Cyclic

Deformation Response and Crack Initiation Mechanisms under Pure Compression

2
Fatigue. Preliminary work has been carried out to verify the possibility of crack nucleation

under pure compression fatigue as well as to establish the necessary condition for crack to

nucleate. As in general in the industrial sectors they have always considered fatigue crack

initiation are due to certain forms of stress concentration, this preliminary investigation was

conducted using notched bars. Some of the experimental details and key results are discussed

briefly in the next section.

1.1.1. Recent Study of Crack Initiation under Pure Compression Fatigue Load

In the author‘s preliminary work, a sample material, SAE 1045 was chosen in both

quench-tempered (HRc: 30) and annealing conditions (HRb: 94). Tensile bars were machined

out of these two materials according to ASTM standard E8M[7]. Various circumferential

notched configurations, namely the U and V shaped notches with different notch depth, were

also designed to examine the effect of different stressed concentrators. For reference, two

smooth bars in the quench-tempered condition were also included as the no stress

concentration samples. Under the loading condition of pure compressive cyclic stresses with

R=20, it was revealed that fatigue crack nucleation is possible under pure compression

provided a stress concentration is present[13]. It was further found that fatigue crack

nucleation occurs when a certain level of von Mises stress is reached at the notch root. For

the quench-tempered steel, this threshold von Mises stress was established to be between -

1.57 of the steel‘s yield strength ( and -2.16 as observed in Figure 1-2. Due to the

presence of a notch, a 3D stress state is developed at the notch root. This gives rise to a

localized von Mises Stress that may inflict a severe localized plastic deformation at the notch

front. Thus, the crack initiation process is observed to be associated with plastic deformation

such as grain rotation as Figure 1-3 demonstrates. As a result, the crack was shown in Figure

1-4 to nucleate close to an approximate 45 degree angle with the loading axis where

maximum shear stress is found. Furthermore, since plastic deformation was a precursor for

crack nucleation, annealed samples of higher ductility also exhibited a better withstanding to

3
crack nucleation under compression fatigue than the quench-tempered if weighted stress is

employed.

Although the above work has demonstrated how stress concentration sites may lead to

crack initiation under pure compression fatigue, these do not reflect the true underlying

micro-mechanisms on how cracks nucleate under fatigue from the material perspective. As is

illustrated in Figure 1-5, if a mini sample of the same material was placed at the stress

concentration site, the mini sample experiences elevated stress due to the stress concentrator.

Nevertheless, it still needs to progress through cyclic deformation before crack nucleates. In

our preliminary work, the results were not complete to show such underlying cyclic

deformation mechanisms from the material‘s perspective. Further work is hence needed to

answer the following questions: 1) what is the general cyclic deformation response of a

material when the applied stress spectrum is purely compressive? 2) Would the material

behave the same way as in the symmetrical fatigue condition? 3) What is the deformation

micro-mechanism when the applied stress spectrum is completely compressive? 4) Is it

always true that pure compressive fatigue load would not introduce fatigue crack when there

is no stress concentration present? 5) If fatigue cracks do form under compressive fatigue,

what is the fatigue condition, namely what is the stress level required? Also, what is the

corresponding mechanism?

4
Figure 1-2: Consideration of von Mises stress vs. number of cycles to crack initiation

(quenched-tempered condition) [13]

Figure 1-3: Illustration of cracking along grain rotation in annealed condition [13]

5
Figure 1-4: Crack nucleation in quench tempered sample: (V-Notch) Kt =6.72, Number of

cycles: 17000; Peak Stress: 90%σyc [13]

Generally, cyclic deformation mechanisms are established through: 1) cyclic stress-

strain responses, 2) cyclic hardening curves and 3) cyclic stress strain curves (CSSC). The

corresponding microstructure observations such as surface observation and dislocation

evolution are also required to determine the cyclic deformation mechanisms. Furthermore,

from the surface observations, crack nucleation may be further determined. Thus, the crack

nucleation mechanisms that were yielded strictly from cyclic deformation may also be

established. Notably from the large quantity of literature reviewed, these aspects of fatigue

have been explored relatively extensive under tension-compression fatigue or pure tension

fatigue conditions. Most of these investigations utilized pure copper such as OFHC Copper in

either monocrystalline [14, 15] or polycrystalline form [16-22] as the subject material.

Results have demonstrated that under these fatigue conditions, copper exhibits a very unique

cyclic deformation response with the formation of corresponding distinctive surface features.

These will be discussed in further detail in Chapter 2, the literature review.

6
Figure 1-5: Schematic illustration of the effect of stress concentration site and the relation of

cyclic deformation behaviour of a material

1.2. Objectives

In order to answer the aforementioned questions, a project of fundamental study is

established with the following objectives:

1. To observe the general cyclic deformation response under pure compression fatigue

and to investigate the deformation micro-mechanisms.

2. To verify if cracks would form and under what conditions cracks would form under

pure compression fatigue

 If crack forms, what are the micro-mechanisms in relation to cyclic

deformation response

3. To analyze and reveal the difference between the above cyclic deformation

responses and micro-mechanisms with those of tension-compression fatigue

As there is large available information on fatigue of OFHC Copper in either single crystalline

or polycrystalline forms in previous studies of over 50 years, the same material, i.e., Oxygen

Free High Conductivity copper (OFHC Cu 101) will be employed for the present

investigation.

7
1.3. Structure of the Thesis

The structure of this dissertation is presented in the following manner. Chapter 2

reviews the cyclic deformation behaviour of OFHC Cu 101 in both monocrystalline and

polycrystalline form. This includes the behaviour of both conditions of symmetrical fatigue

and the conditions that are biased with non-zero mean stress. The general fatigue crack

nucleation mechanisms will also be discussed including fatigue crack behaviour under pure

compression fatigue. Furthermore, the asymmetrical behaviours, the Bauschinger effect and

the negative Bauschinger effects will also be presented. Chapter 3 consists of the

experimental strategies, set-up and procedures that were conducted in this project. Results

and discussions that derived from these experiments are presented in the subsequent chapters.

Chapter 4 focuses on the cyclic stress strain response and microstructure evolution of

polycrystalline Cu under pure compression fatigue. In this chapter the fundamentals of cyclic

deformations under pure compression are established. These include the cyclic stress-strain

response, surface morphology evolution and the dislocation evolution under pure

compression fatigue. Chapter 5 presents the results and discussion on the exploration of

crack nucleation under pure compression fatigue. The chapter provides the determination on

the possibility of crack nucleation as well as the micromechanisms behind such nucleation

process. Chapter 6 discusses the results of asymmetrical mechanical response of the OFHC

polycrystalline copper. The phenomenon was observed to occur both in uniaxial mechanical

testing as well in the cyclic testing. Finally, the entire dissertation will be summarized and

concluded in Chapter 7 with Chapter 8 considering recommendations for future work.

8
2. Literature Review

2.1. Overview

Fatigue behaviour, in general practice, is studied through the establishment of

Wholer‘s curve [23] or S-N plot as depicted in Figure 2-1. By plotting the applied stress

against the log number of cycles required to fracture each of the samples, the fatigue

resistance of the material or the fatigue endurance limits is obtained.

Fatigue Endurance Limit


Stress (MPa)

Log (No. of cycles)

Figure 2-1: Illustration of fatigue endurance limit

It is interesting to note that fatigue endurance limit of a material is generally well-

below its own yield strength or rupture strength. In other words, it was seen that even if the

fatigue strength is well below the material‘s yield or rupture strength, fatigue damage or

cyclic deformation still occurs. For many years, this phenomenon had posed a great question

to the scientists and experts in the related field. It was subsequently found that cyclic plastic

straining was the main cause for the fatigue failure. Despite the fatigue strength being lower

compared to the materials‘ yield or rupture strength, the stresses are large enough to

introduce the required microplastic strains, which are typically in the range of 10-5 to 10-4.

9
Consequently, fatigue failure is considered as the accumulation of such plastic straining. As

the S-N curve determines only the materials‘ total fatigue lives, it certainly does not answer

how the fatigue ―damage‖ originates from the cyclic deformation process nor does it describe

cyclic plastic straining development. The investigation of cyclic deformation mechanisms

and the explanation to the nature of fatigue damage are therefore needed.

Traditionally, cyclic deformation mechanisms are studied through the combination of

mechanical stress-strain responses and the evaluation of the microstructure evolution that

occurs during the cycling process [24]. These studies, for instance, deal with the examination

of the surface morphology evolution and the changes in dislocation structures upon cycling

with the corresponding stress-strain responses established. As stated previously in Chapter 1,

OFHC copper is selected as the material of interest on the basis that cyclic deformation of

copper is well established. In this chapter, some fundamentals of cyclic deformation response

and mechanisms of monocrystalline and polycrystalline of Cu are discussed. To address

these topics the cyclic deformation behavior of copper under symmetrical fatigue is first

described in section 2.3. Subsequently, section 2.4 is a discussion on the fatigue behavior

under the influence of mean stress, i.e., the effect of mean stress. Although crack nucleation

does not always accompany cyclic deformation, in the event that cyclic deformation does

lead to crack nucleation, these underlying mechanisms need to be considered. Section 2.5

identifies some of the major crack nucleation mechanisms in general, where section 2.6

describes fatigue crack behaviour under pure compression fatigue in particular. Furthermore,

as cyclic testing is essentially a repeat of forward and reverse loading process, Bauschinger

effect should be revisited as presented in section 2.7. Several terminologies of cyclic

deformation response are defined in section 2.2.

10
2.2. Terminology of Cyclic Deformation Response

As a standard, cyclic testing is conducted in either a controlled closed-loop stress or a

controlled closed-loop strain condition, and the obtained respective response is therefore

constructed in the form of stress-strain hysteresis loop. A hysteresis loop reveals basic but

key information on cyclic stress-strain behaviour of a material. For example, the loop width

would signify the amount of plastic strain accommodated by the material per cycle. Since

fatigue failure is a result of cyclic plastic straining, this information would be particularly

significant. Consequently, the cumulative plastic strain is used to assess and describe the

damage accumulation between different fatigue testing conditions. Furthermore, depending

on the material as well as loading condition, there may be changes in the hysteresis loop

behaviour such as the loop width change or the shift of loops throughout the cyclic

deformation. These behaviours are further classified into the following two cyclic stress

deformation responses: cyclic softening/cyclic hardening and cyclic creep. In the next

sections, these terms will be introduced and characterized.

2.2.1. Cyclic Softening/Hardening Behaviour

The phenomenon of cyclic softening is defined when the observed hysteresis loop

width becomes wider in loop width in the stress controlled conditions or as such the stress

response of the hysteresis loop becomes smaller in the strain controlled conditions. Cyclic

softening reflects the increase in material‘s ability to accommodate plastic strain under the

constant stress. Schematics of these responses are shown in Figure 2-2 where (a) represents

stress-controlled and (b) total strain controlled compression-compression conditions. For

cyclic hardening, the sequence of cyclic deformation response is simply in the reversed order

as shown in Figure 2-2. The phenomenon of cyclic softening or cyclic hardening behaviour is

the reflection of plastic strain accommodation, which results in the microstructural changes.

Thus, these behaviours are generally transient behaviours and would eventually come to a

saturation, after which continuing cycling would not result in any changes in stress (strain

11
controlled test) or strain (stress controlled test). Conventionally, it is a general trend that the

well annealed materials, which have low dislocation density should cyclic harden first before

reaching to saturation, whereas a heavily worked material with high dislocation density

would cyclic soften prior to saturation. [25]

Figure 2-2: Illustration of cyclic softening behaviour under (a) stress controlled, and (b) total

strain controlled compression-compression tests. The reversed trend is observed

for the cyclic hardening behaviour.

2.2.2. Cyclic Creep

In the case of cyclic creep, a shift of the hysteresis loop along the strain axis with time

is observed in a constant stress amplitude controlled test as illustrated in Figure 2-3 (a). The

occurrence of cyclic creep arises when there is unequal plastic deformation accommodation

between the forward and reverse load directions [26]. The phenomenon is often observed

when the applied load spectrum is offset from the symmetrical conditions, namely when the

mean stress is not zero. Conversely, in a constant strain amplitude controlled test, similar

phenomenon can be observed for which there is a tendency for mean stress moving towards

zero as shown in Figure 2-3 (b). Such behaviour is known as stress relaxation. Analogous to

12
cyclic creep for stress-controlled tests, non-zero mean strain loading condition is required for

the occurrence of stress relaxation.

In addition, as plastic strain is measured by the hysteresis loop width, the cyclic creep

curves show the accumulation of permanent deformation encompassing two processes: (i)

increase in the cyclic plastic strain, i.e., the shift in mean strain, and (ii) the cyclic plastic

strain accommodation. The method of evaluation of these strains, and , on

hysteresis loop is therefore adopted from Gaudin and Feaugas as illustrated in Figure 2-3 (a)

[27]

Figure 2-3: Schematic of cyclic creep behaviour under (a) stress controlled, and (b) total

strain controlled compression-compression tests.

2.3. Cyclic Deformation Behaviour of Copper under Symmetrical Fatigue

For over the last half century, cyclic deformation behaviour of Cu, in both single

crystalline [14, 15] and polycrystalline forms [16-22], has been studied extensively. During

the beginning of this period, much of systematic work was carried out to study the cyclic

stress-strain response of single crystal Cu oriented for single slip under symmetrical fatigue.

The rationale for the use of monocrystalline samples was 1) easy manipulation of the

crystal‘s orientation and 2) easy determination of the relationship between the orientation and

13
the loading condition such as the slip geometry and the resolved shear stress [24]. The results

yielded from these investigations are then implemented as the framework for studying the

cyclic deformation behaviour of polycrystalline Cu. Detailed summaries for both single [28]

and polycrystalline [29, 30] Cu have also been published following these investigations. The

focus of this review will first consider monocrystalline Cu.

2.3.1. Monocrystalline Copper

It was determined that under symmetrical fatigue conditions, irrespective of load or

strain control, well annealed single crystal copper with easy slip orientation experienced

cyclic hardening until the eventual saturation behaviour is reached. Such saturation is often

summarized in the form of a cyclic stress-strain curve (CSSC). Unlike the establishment of a

tensile stress-strain curve, which is obtained in one static tensile test, the cyclic stress strain

curve is established through numerous cyclic tests at different peak stress magnitudes (in the

case of stress controlled cyclic test). In other words, each of the data points on a cyclic stress

strain curve represents the pair of designated peak stress magnitude and the corresponding

strain at saturation. By the same analogy, for strain controlled cyclic tests, each of the data

set on cyclic stress strain curve is taken as the pair of given strain and its corresponding

―saturation‖ stress. Once the CSSC is determined through a wide range of cyclic testing

conditions, it is observed that there are three distinctive regions on the established CSSC, as

seen from Figure 2-4: Region A, the rapid cyclic hardening rate region and Region B, an

intermediate plateau followed by Region C, another rapid increase in saturation strain region.

As with all explanations of metallic materials‘ mechanical deformation response, the CSSC

behaviour is closely linked to the microstructural changes in the material, for instance, the

surface morphology and/or dislocation evolution. The cyclic stress-strain response of

monocrystalline Cu is of no exception. It was found that this behaviour was closely linked

with the dislocation arrangements. Each region is discussed individually hereafter.

14
Figure 2-4: Cyclic stress strain curve [14]

Region A marks the low plastic shear strain amplitude region, (γap < 6 × 10-5). In this

region cyclic hardening is mainly due to the accumulation of the primary dislocations which

are edge in character. Dipoles are formed and agglomerated into dislocation bundles as

shown in TEM micrographs (Figure 2-5 (a)). The saturation state in this region is the result

of the balance between the dislocation-rich bundles and the dislocation-deficient areas. It was

proposed that the deformation within the matrix vein is accommodated by the flip-flop

motion of dislocation loops, which are produced by jogs during cross-slip of screw

dislocations [31, 32]. Hence, only fine slip markings are observed on the surface of the

sample. It is said that if a material is cycled within this region, it has almost infinite fatigue

life. Namely, cracks do not nucleate in Region A.

15
Figure 2-5: Three-dimensional TEM micrograph of the dislocation arrangement in

monocrystalline Cu in different cyclic saturation region (a) Region A: γap = 2.6 ×

10-5 (b) Region B: γap = 1.5 × 10-3 (c) Region C: γap = 1.45 × 10-2

Region B denotes the plateau regime of CSSC curve, where the plastic strain is

independent of stress (under stress controlled condition or vice versa under the strain

controlled condition) [15]. Of the three regions, Region B has attracted the most interest, as it

indicates the formation of persistent slip bands (PSBs), a microstructure reflecting uniquely

the materials‘ cyclic deformation response. With its distinctive ladder-like dislocation

structure (as shown in Figure 2-5(b)) that is composed of heavy dipolar walls, PSBs were

known to accommodate the majority of the cyclic plastic strain leading to the phenomenon of

strain localization. As a matter of fact, it was documented by Winter [33] that the plastic

deformation is so concentrated in PSBs, the PSBs deforms about 100 times more than the

matrix. The heterogeneous deformation between the matrix and the PSB is also described by

Winter‘s rule [33] as shown in Eq. 2-1,

γap= f PSB × γap,PSB +(1- f PSB ) × γap,M Eq. 2-1

16
where fPSB is the volume fraction of PSBs, γap,PSB, γap,M are the local plastic shear strain

amplitude acting in the PSBs and matrix respectively. While the PSBs are supporting most

of the plastic shear strain, they experience macroyielding, which involve dislocation

multiplication. Edge dislocations are bowing-out from walls and transport along the channels.

Screw dislocation in the channels may also draw the edge dislocations out of the walls as

demonstrated schematically in Figure 2-6. Figure 2-7 is a TEM micrograph corresponding to

this phenomenon, where there is an apparent primary edge dislocation bowing out of the

walls.

Figure 2-6: Dislocation

arrangements in

persistent slip bands in

FCC metals [34]

Figure 2-7: TEM of a section parallel to the primary slip plane of single crystal Cu fatigue to

saturation at γap = 5 × 10-3 [35]

17
In general it was found that at low shear strain amplitude the dislocations in matrix

display mainly bundle and vein structures. Similar to Region A, only single slip prevails at

low shear strain amplitude in Region B with a ladder-like dislocation structure. As the shear

strain increases, secondary slip gradually gains importance. This leads to the formation of

labyrinth dislocation structures at high shear strain amplitude of Region B. Furthermore,

according to Blochwitz and Veit [36], who tried to determine the ―true CSS curves‖ and

hence the τs required to nucleate PSBs, they had established that at τs the matrix has the

corresponding stress amplitude required to initiate the PSBs. It is, in other words, at this

constant τs the increase in γap can be deemed as the triggering mechanism for the nucleation

of PSBs. Thus, the appearance of the plateau is explained. As such, τs is a constant and the

intensification of γap is due to the increase in PSB volume fractions. Eventually, extrusions

and intrusions are merged on the surface as the surface manifestations of these PSBs. Since

these PSBs are generally considered as the very mechanism for fatigue crack initiation, the

detailed formation of PSBs is described in section 2.5 where the general crack initiation

mechanisms are discussed.

Moreover, the plateau stress for monocrystalline Cu is reported as τs = 28MPa under

strain controlled conditions and τs = 32MPa under stress controlled conditions [14]. Similar

values were also documented elsewhere [35]. However, this plateau stress is dependent on

testing conditions such as temperature for the reason that dislocation arrangements are

sensitive to temperature. It was found PSBs could not exist at high temperature and at low

temperature extended dislocation wall structures are observed instead. This invariably

signifies the plateau signature for monocrystalline Cu is merely the reflection of PSB

formation for Region B or as such that PSBs determines the τs values.

Region C describes high plastic shear strain amplitude region, (γap > 7.5 × 10-3).

Secondary slip plays an important role in this region leading to the formation of dislocation

18
cell structures, as seen in Figure 2-5 (c). Ladder structure of PSBs also cease to exist and

instead equiaxed cell structures are presented. Unlike the occurrence of localized plastic

deformation in Region B, the deformation in Region C is completely homogenous within the

bulk of the specimen gauge length.

Overall, the shape of CSSC is relatively unaffected by the orientation of the single

crystal. For most single-slip orientations, τs values follow those presented in Figure 2-4. In

contrast, for monocrystalline Cu that is oriented for multi-slip, a very dissimilar behaviour is

seen [37-40]. As demonstrated by Gong et al. [39] and Wang et. al. [40], when

monocrystalline copper was oriented in multi-slip direction, i.e., [100] direction, the

existence of the plateau region may be completely eliminated. The resulting dislocation

configuration was also found to be a labyrinth structure instead of the typical ladder like

structure that is generally favoured in the single-slip orientation. These behaviours of

monocrystalline Cu may be representative to the behaviour of individual grains in

polycrystalline Cu [24]. However, whether the grains in polycrystalline Cu behave like that

of monocrystalline Cu collectively as a whole may need further consideration. Namely, if the

same information is also applicable to polycrystalline Cu such as the occurrence of PSB

formation and if a CSSC of polycrystalline Cu may also show the same three distinctive

regions. Hence it leads to the next section.

2.3.2. Polycrystalline Copper

Following the establishment of the cyclic deformation response of single crystals,

many also investigated the cyclic stress-strain response and the respective CSSC of

polycrystalline copper under symmetrical fatigue condition [16-22]. It was concluded that

fully annealed polycrystalline Cu, as its single crystal counterpart, also displayed the cyclic

hardening and saturation behaviours under symmetrical fatigue. However, the results on

whether the CSSC of polycrystalline Cu also exhibits the same type of plateau were

19
inconsistent. It was found that the existence of plateau behaviour may vary depending on

multiple factors, such as grain size [41-45] and testing conditions [16, 22, 46], thus making

the correlation of monocrystalline data to polycrystalline difficult.

By nature, polycrystalline material is composed of many monocrystals of the same

material that are oriented randomly with respect to each other. This provided the physical

basis for researchers and scientists to develop and employ the composite models. Typically

three composite models are used to relate the results that are obtained from monocrystalline

to that of the polycrystalline. They are Taylor, Sachs and Schmid [24], respectively, with

common expressions as shown by Eq. 2-2 and Eq. 2-3

Eq. 2-2

Eq. 2-3

where and are the values of shear stress and shear strain that are obtained from the

tests with monocrystalline specimen, and and are the predicted values for stress and

strain for polycrystalline specimens. M is the orientation factor and differs in value between

the 3 models.

For Schmid‘s Model, the assumption is that individual grains are oriented in the

direction favourable to easy glide or single slip and the compatibility issue between

individual grains is ignored. As a result, M is equal to 2. Sachs model, on the other hand, also

assumed that there is no contact interaction between individual grains but each grain is

oriented at random slip directions. It is also further assumed that each grain also experiences

the equivalent deformation strain. Hence M has the value of 2.24. The Taylor model, in

contrast to the previous two models, has taken into the consideration of the contact between

grains by considering them fully physically connected. Like Sachs model, the grains have

20
random orientations but have the same strain. This model is hence deemed as the most well-

known model of the three with M = 3.06.

It was found that for testing conditions close to the current research, i.e., under load

control and with similar grain size, the CSSC of polycrystalline Cu also exhibits three

distinctive regions. However, instead of the true plateau behaviour, it shows quasi plateau

behaviour [19, 47]. Specifically, Wang and Laird [19] had determined that this plateau stress

corresponds to 98MPa, which is equivalent to the plateau stress predicted by Taylor‘s model

(i.e., 3.06 32 = 98MPa for stress control test) as shown in Figure 2-8. It was also concluded

from their work that ordinary strain control does not give the appearance of the plateau. It

only occurred when the material is initially ramp loaded. Although Figueroa et al, [16] did

not apply ramp loading, they also found the quasi plateau tends to be more obvious in stress-

controlled fatigue tests than that of strain-controlled testing condition as illustrated in Figure

2-9. Nevertheless, it was further revealed that the quasi plateau is also associated with the

formation of PSBs [16-22]. Figueroa et al, had found that in load-control symmetrical

fatigue, despite not as regular, similar ladder-like PSBs structures could also be seen in the

polycrystalline Cu as seen in Figure 2-10 [19] and these ladder-like structures are formed by

the dipolar wall structures [16, 21].

21
Figure 2-8: Comparison of the

cyclic stress-strain

curves. The shear

stress is transposed

by the Taylor

factor [19].

[48]

[49]

[50]&[51]

[52]

Figure 2-9: Cyclic stress-strain curve of annealed copper shown in semilogarithmic plot

adopted from [16]

22
Figure 2-10: Typical ladder-like PSBs in bulk of polycrystalline: (a) straddling a twin

boundary; (b) contained entirely within a narrow twin; specimen ramp-loaded to

98MPa, and then step-tested in strain control to the midpoint of the plateau at

98MPa [19]

Evidently loading conditions have certain influences on the shape of the CSSC as

seen from the difference between load-control vs. strain-control in Figure 2-9. As mentioned

previously, multiple factors could dictate the shape and the position of the CSSC. These

include the experimental conditions such as the start-up loading condition and the application

of mean stress [19, 22]. To briefly summarize the effect of start-up condition, when changing

the loading condition from zero to ramp loading under symmetrical fatigue condition in

stress control, it was found that depending on the length of the ramp length, the magnitude of

cyclic saturated plastic strain may be different. As such, when the ramp length is increased,

the CSSC is shifted towards smaller values of saturated plastic strain[22]. Since the testing

condition in this project does not involve ramp loading and is subject to the pure compressive

23
loading spectrum, the consideration of mean stress effect is thus essential. Literature review

on the effect of mean stress was therefore conducted and is presented in the following

sections.

2.4. Cyclic Deformation Behaviour under the Influence of Mean Stress

The understanding of mean stress effect on fatigue is especially important since

fatigue conditions in practical applications are not all symmetrical. Much like fatigue

behaviour in general, the effects of mean stress on the fatigue behaviour of materials are

commonly explored through its relationship with material‘s fatigue life and fatigue strength.

i.e., how the mean stress alters the S-N curve or fatigue‘s strength of fatigue life [53-64].

Often, these works focus on determining empirical relationships for predicting fatigue lives

and strength under mean stress effects or the combined effect of mean stress and stress

amplitude [53-55, 57, 59] or redefine the known relationships for different materials[57, 58,

61, 62, 65, 66]. Generally, the application of mean stress gives rise to cyclic creep,

frequently all called ratcheting, under stress-controlled tests [58, 62, 64, 65, 67-69]. Several

constitutive models were also developed to describe the relationship between the effect of

mean stress and stress amplitude on the ratcheting behaviour and subsequently relating the

established ratcheting behaviour to fatigue life [60, 63, 64, 68-71]. Comparing the number of

studies of mean stress effect on fatigue life, strength, and the ratcheting behaviour, the effect

of mean stress on cyclic plasticity [67, 72-75] is much less reported. As fatigue damage is

mainly derived from cyclic plasticity, the effect of mean stress on cyclic plasticity is

therefore considered first in the next section followed by the effect of mean stress on fatigue

strength and life.

2.4.1. Effect of Mean Stress on Cyclic Plasticity

The influence of mean stress on cyclic stress-strain (CSS) response and the

corresponding CSSC is often studied under positive mean stress conditions. It was found that

24
under the application of mean stress the corresponding saturated plastic strain amplitude may

be altered. For some it was suggested that there is a strong influence of tensile mean stress

on the saturated plastic strain amplitude as such the CSSCs for symmetrical fatigue is

considerably different than those obtained for pulsating tension fatigue. These include results

acquired by Lorenzo and Laird [57] and Eckert et al. [58] under high-amplitude loading

conditions as well as Lukas and Kunz [67] for low-amplitude cycling and both for

polycrystalline copper. In addition, Lorenzo and Laird [57] and Eckert et al. [58] had further

developed a linear relationship between log of plastic strain range ( ) on mean stress

( ) under a constant peak stress ( ) , where the dependence of log on

exhibits a negative slope. A set of parallel family of lines were observed for different

magnitude.

Nonetheless, the overall trend on the effects of mean stress on cyclic plasticity is

found to be rather diverse. In some cases [16, 56], it was found that the saturated plastic

strain amplitude is unaffected by the mean stress. For examples, the results from Manson and

Halford‘s on steels [56], Lorenzo and Laird [76] on monocrystalline Cu and Figueroa et al.

[16] on polycrystalline Cu at low stress amplitude. There are also cases where the saturated

plastic strain amplitude increases with increasing tensile mean stress such as the observation

made by Pokulda and Stan [72] for steels under low constant stress amplitudes.

On the other hand, there are other investigations that had observed decreases of the

saturated plastic strain amplitude with the increase in tensile mean stress under the same

stress amplitude [58, 67, 74]. Namely, the CSSC shifts to lower plastic strain amplitude as

the mean stress increases. This includes the results for steels obtained by Pokulda and Stan

[72] under high constant stress amplitudes. Turner and Martin [77] also observed the same

trend for Type 304 stainless steels under constant peak stress magnitude but at high stress

25
amplitudes. The same findings were obtained by Kliman and Bílý [73] on carbon steel

irrespective of the magnitude of the stress amplitude.

A more recent and interesting observation made by Lukas et al. [75] is that, regardless

of the effect of mean stress on its saturated plastic strain amplitude, the typical dislocation

arrangements under the bias of positive mean stress is found to be cell structures without any

ladder-like PSBs in monocrystalline copper as shown in Figure 2-11. Eckert et al. [58] also

noted that cellular dislocation structure was observed in their result on polycrystalline

copper. Correspondingly, Lukas et al [75] had found that, instead of the persistent slip band

formation, coarse slip bands formed on single crystalline copper which was cycled under

pulsating tension fatigue as indicated in Figure 2-12. They had developed a stress

requirement map for which conditions PSBs are able to form as show Figure 2-13. It is

observed that, PSB formation occurs at the condition where fatigue fracture is expected and

extending below the fatigue limit.

Figure 2-11: Dislocation cell structure

observed in monocrystalline Cu

under pulsating tension fatigue

[75]

26
Figure 2-12: Coarse slip bands on surface of

monocrystalline Cu under

pulsating tension fatigue [75]

Figure 2-13: Stress requirements for the

formation of PSBs in single-slip-

oriented copper single crystals

(schematic)

2.4.2. Effect of Mean Stress on Fatigue Strength and Fatigue Life

Figure 2-14 (a) shows the generic effect of mean stress on the S-N curve. When the

stress amplitude (σa) is plotted against the number of cycles (N) in log scale with respect to

different magnitudes of mean stress (σmean), the fatigue life is observed to decrease with the

increase in mean stress at a given σa. On the same plot, fatigue strength is also seen to be

significantly lowered with the increased in mean stress.

27
Figure 2-14: (a) Typical stress amplitude-life plots for different mean stress values adopted

from Hertzberg [78] (b) Gerber, Goodman, and Soderberg diagrams (R-M

diagram) showing combined effect of alternating stress amplitude and mean

stress on fatigue endurance [53-55]

The effect of mean stress on fatigue could also be represented using a constant-life

diagram, Figure 2-14 (b), which is also known as the R-M plot where the R stands for range

of stress and the M stands for mean stress. The constant-life diagram, Figure 2-14 (b), shows

the three widely used relations, Gerber [53], Goodman[54] and Soderberg[55] relations and

are described mathematically in Eq. 2-4, Eq. 2-5 and Eq. 2-6 respectively. Depending on the

mean stress (σmean) magnitude, these equations could be used to calculate the maximum

allowable stress amplitude that the material could have before fatigue failure based on the

material‘s fatigue strength (σfat), and/or tensile strength (σts) and yield strength (σy).

Subsequently, the three lines on Figure 2-14 (b) also mark the boundary between the safe and

unsafe design criterion. As evidence suggests [26, 59], most experimental data shows that

28
theoretical conditions fall between the Geber and Goodman lines, Soderberg relation hence

gives the most conservative estimate.

Goodman Relation: ( ) Eq. 2-4

Gerber Relation: * ( ) + Eq. 2-5

Soderberg Relation: ( ) Eq. 2-6

Analogous to methods of approximating fatigue lives in general, to estimate the effect

of mean stress on fatigue lives, two more alternative approaches are devised: stress-life and

strain-life approaches. For the stress-life approach, Morrow has proposed following equation,

Eq. 2-8, based on Basquin‘s fatigue relation (Eq. 2-7), where , are the fatigue

strength coefficient and exponent as well as the fatigue life respectively from the Coffin-

Manson equation. According to Morrow and Eq. 2-8, fatigue strength coefficient is reduced

with increase in tensile mean stress where the compressive mean stress increases it.

Additionally, he had developed a strain-life approach, Eq. 2-9. Similar to Coffin-Mason

relationship, , and are the respective terms for total strain range, fatigue ductility

coefficient and exponent.

Basquin‘s Fatigue Relation [79]: Eq. 2-7

Stress-life Approach[80]: Eq. 2-8

Strain-life ⁄
( ) ( ) Eq. 2-9
Approach[80]:

29
In some investigations under low cycle fatigue [62, 66, 68], it is seen that Smith,

Watson and Topper (SWT) [81] as well as Walker [82] relations are used to express the

effect of mean stress. These relationship are [62, 66]:

SWT [81]: √ Eq. 2-10

Walker [82]: Eq. 2-11

where in Walker‘s equation (Eq. 2-11) is a material constant. It was seen that for some

materials such as Elbrodur-NIB copper alloys, the SWT criterion gives a better estimate than

the Walker‘s model. For other materials such as some aluminum alloys, Inconel 718 and

titanium alloy, Walker criterion yields a better trend than SWT[62, 66].

[83]

[84]

[85]

Figure 2-15: R-M diagram showing the effect of tension and compression mean stress

adopted from Forrest [86]

30
Overall, irrespective of the method of approach, it was found that in general the

existence of tensile mean stress reduces the stress amplitude that a material can withstand

while the addition of compressive mean stress may lengthen it [59, 86-88]. Figure 2-15 is a

summary of data combined from three different sources [83-85] that suggests this trend when

the ratio of stress amplitude to the fatigue strength is plotted against the ratio of mean stress

to the yield strength [86].

Figure 2-16: Effect of mean-stress on fracture mechanism in stress-controlled tests [88]

Moreover, the effect of mean stress on fracture mechanisms in stress-controlled

fatigue is summarized schematically by Klesnil and Lukas in Figure 2-16 [88]. Figure 2-16 is

divided into four fields based on different failure mechanisms [88]. For stress amplitudes

below the fatigue limit line, the effect of mean stress on fatigue life is negligible. Namely, no

fatigue failure is expected irrespective of the mean stress magnitude. For stress amplitude

higher than the fatigue limit, it is observed that the diagram is comprised of three regimes.

The middle region classifies the conditions where fatigue failure is found. It is bounded

between the two regimes at each end, where cyclic creep failure mechanisms dominate.

Incidentally in the high tensile mean stress regime, it is expected that final failure should

31
follow that of the regular uniaxial test, i.e., by necking, the ductile type of failure, whereas in

the high compressive mean stresses region, buckling would be the likely outcome due to the

―loss of stability‖ [88]. It was also commented that in these regions, where cyclic creep

failure dominate, there is strong of effect of R ratio, where R= σmin / σmax [58].

2.5. General Fatigue Crack Nucleation Mechanisms

Cyclic deformation does not necessarily yield crack nucleation but crack nucleation is

the result of cyclic deformation. Depending on the magnitude cyclic plastic strain

accommodation or the severity of cyclic damage, crack nucleation may arise from cyclic

deformation. When a crack initiates during cyclic deformation, it was found that it is almost

always nucleates at free surface of the component [78, 89-91]. Both Wohler [23] and

Bullens [92] had shown that serious reduction in fatigue life is associated with the loss of the

surface quality. Hence, from an engineering applications perspective, fatigue crack

nucleation mechanism is often thought to be governed by certain forms of stress

concentrators. These may be:

i. Surface roughness (incursion during manufacturing process) or notches

ii. Surface protrusions due to the formation of PSBs [93]

Moreover, from series of experiments conducted by Alden and Backofen [94] and

Mughrabi [95], if the specimens that had fatigued substantially relative to the fatigue life

were refurbished to their original surface state, they showed that their fatigue lives were also

restored to that of the virgin smooth samples. Therefore, it was concluded that major

emphasis should be placed on surface state when it came to studying fatigue life. As fatigue

strength is significantly reduced by these stress concentrations, the effect of stress

concentrators is discussed first. Then, the formation and role of persistent slip bands (PSBs)

in fatigue will be covered. Finally, the effect of grain and twin boundaries will be considered.

32
2.5.1. Effect of Stress Concentrators

The impact of stress concentrators on fatigue life is generally investigated through

notched specimens that contain either a V notch or a circular notch [64-66]. These surface

stress concentrators or surface discontinuities introduce a non-uniform and complex stress

field in the adjacent surroundings by bringing in the three effects: 1) the local peak stress at

the root of irregularity is much greater than the average or nominal stress across the section,

2) a pronounceable stress gradient is produced and 3) the development of a triaxial stress

state [26, 96]. Depending on the geometrical dimension of the surface irregularity, either in

shape (i.e. sharpness of notches) or the depth, it will also give divergent effects on the fatigue

strength [97].

To assess the effect of notches on fatigue strength, a comparison of fatigue endurance

limits between the notched and un-notched specimens is made and is typically expressed by

the fatigue notch factor, Kf. Kf is simply the ratio of the fatigue limit of un-notched to that of

notched specimens.

 end , unnotched
Kf  Eq. 2-12
 end , notched

Values of Kf could also be altered by factors like 1) stress level, 2) loading method, 3) notch

type, 4) the severity of the notch, and 5) the material [26] . Once Kf is established, the fatigue

notch-sensitivity-factor, q, may also be computed, where Kt is the stress concentration factor

in relation to the applied load.

Kf 1
q Eq. 2-13
K t 1

Incidentally, q has values ranging from 0 to 1. Under q = 0 (i.e., Kf =1), the material

experiences no effect from a notch. For q = 1 (i.e., Kf = Kt), notch has imposed full effect on

33
the material. Equation 2-12 gives a simple indication on how effective the notch is in

reducing the material‘s fatigue strength [2].

The above has demonstrated how surface roughness may lead to crack initiation.

However, this does not reflect the true underlying micro-mechanisms on how cracks nucleate

under fatigue from the material perspective. It certainly does not explain how fatigue cracks

form at smooth free surfaces of metals, alloys and commercial materials, either. Thus the

attention is immediately directed to fatigue crack nucleation micro-mechanisms in nominally

defect-free and un-notched materials.

2.5.2. Role of Persistent Slip Bands & Intrusions and Extrusions

Investigations have shown that crack initiation mechanism of a defect-free and un-

notched pure OFHC copper has always been attributed to the formation of PSBs at the free

surface [26]. In 1953, Forsyth made a first documentation on the observation of extrusion

from the PSBs in a solution-treated Al-4wt% Cu Alloy and again in 1957 he observed the

same in both single crystals and polycrystals of silver chloride [98, 99]. Later, both Forsyth et

al [100] and Cottrel et al.[101] showed that extrusion formation comes in pair with intrusion

on copper samples during fatigue. A picture of typical extrusions and intrusions on a fatigued

surface of monocrystalline copper is shown in Figure 2-17 [102]. The rise of sharp peaks and

troughs, which result from PSBs, is the most unique characteristic observed in fatigued

materials [78].

34
Figure 2-17: Some PSB protrusions with superimposed extrusions and intrusions in a copper

single crystal. A crystal tested at a plastic strain amplitude of 2 x 10-3 for

120,000cycles [102].

Generally, fatigue crack nucleation mechanism in metals and alloys of high purity has

often been associated with Wood‘s postulate made in 1958 [26, 103]. He proposed that

different irreversible net slips are created on different slip planes in the material under cyclic

straining. They appear on the free surface as microscopic ‗hills‘ and ‗valleys‘, i.e. protrusions

and intrusions, respectively, and roughen the surface of materials. Subsequently, these

intrusions would act as micro-notches that increase stress concentration at the roots of these

locations and hence promote further slip and crack initiation such as demonstrated in Figure

2-18 [26, 103]. In Figure 2-18, a simplified model is proposed of which it is assumed that

there is only one slip system operating and that their slip motion is considered as the ―relative

motion of parallel cards‖.

35
Figure 2-18: Model of card slip in fatigue slip band [24].

Mott [104] added to Wood‘s speculation while attempting to explain the mechanics

behind the development of net slip offsets. The to-and-fro motion in fatigue has enabled

screw dislocations to glide along different paths in slip bands by cross slipping, leading the

screw dislocations to complete a circuit during a fatigue cycle. This creates a displacement

on the surface as extrusion with magnitudes proportional to the Burgers vector. Although,

Mott‘s hypothesis of shear displacement on free surface has not been fully validated through

experiments, it serves as basis for subsequent models. For further implementation Kennedy

[105] suggested the gating mechanisms that justified how the screw dislocation motions turn

into irreversible displacements in the forward-reverse oscillations. These gating mechanisms

are 1) obstacles formed to prohibit the dislocation motion, for example, creation of jogs as a

result of edge-screw intersections and 2) the complex interaction between two screw

dislocations and a third dislocation at a node on a free surface.

In short, PSB formation is an outcome that is given by material‘s ability to cross slip.

It also satisfies the following three conditions [106-108]:

i) There is a high localized cyclic strain in this zone.

ii) It has different dislocation structure than the surrounding matrix. It is observed that

36
within PSBs, as shown in Figure 2-19, dislocations structures exhibit a ladder-like

wall configuration, whereas, the surrounding matrix has vein structure [106].

iii) The zone ends on the specimen surface resulting in intrusions and extrusion in

either single crystal or a single grain of a polycrystal.

Figure 2-19 : Ladder like structures and stage I cracks within them [106]

Figure 2-20: Schematic representation of the mechanism of extrusions/protrusions at the PSB

according to Essmann et al. [109]

37
The surface roughening, i.e. PSB extrusions and intrusions, on the other hand, is the

product of systematic build-up of fine slip movements that emerge at the free surface [2,

110]. According to Essmann et al.‘s model [109], the cross-slipping in PSBs increase

dislocation density within the bands. Therefore, they enable the annihilation of the

dislocations with opposite sign thereby creating either vacancies or interstitials within the

PSBs, which lead to extrusions or intrusions respectively at the surface of the component.

Figure 2-20 displays a schematic representation of mechanisms of extrusions.

Neumann [111] has also proposed a model for formation of cracks by coarse slip in

which duplex slip is involved. Figure 2-21 shows cracks are developed through a series of

coarse slip steps. In tension, slip occurs on plane 1, Figure 2-21 (a). A step created by the slip

act as the stress riser which leads to the activation of slip plane 2 under the same tension load,

Figure 2-21 (b). Under compression, all the steps are closed up and ―A‖ serves as the micro-

crack nucleus, Figure 2-21 (c). As the cycles continues, the same repeated slip movement

prompts further microcrack development through the same motion of Figure 2-21 (a) to (c)

and hence leading to crack nucleation. It should be pointed out that the average plane

proposed for this model is perpendicular to the maximum tensile stress as opposed to the

model of intrusion formation on the maximum shear plane in Figure 2-18.

38
Figure 2-21: Neumann‘s model of crack nucleation. In part (c) A represents a crack nucleus

[111]

Overall, after PSB intrusions and extrusions are formed, they would act as miniature

stress risers that provide geometrical stress concentrations. With localized high cyclic plastic

strain at the root of intrusions, it would allow the crack to yield at these sites. Moreover, due

39
to the difference in dislocation density relative to the matrix, it creates a gradient in strain

between PSBs and the bulk that further leads to cracking at the plank of PSBs [78]. Hence,

PSBs in fatigue are said to be prone to crack initiations and is the basic type of crack

nucleation in fatigue [2, 88, 108]. Figure 2-22 a), demonstrates a crack initiating at PSBs

intrusion where Figure 2-22 b) illustrates cracking at PSB-matrix interface [112, 113].

Figure 2-22: a) A section through a PSB containing intrusions at A and B. Note the crack has

started at B [112] b) Fatigue crack initiation (denoted by the arrow) at PSB-

matrix interface in a Cu crystal fatigue for 60,000 Cycles at γpl = 0.002 at 20°C

[113]

2.5.3. Effect of Grain and Twin Boundaries

One of the major differences governing the cracking mechanism in single crystal and

polycrystalline materials is the existence of grains and hence the grain boundaries. Grain

boundary cracking is also a result of cyclic slip processes. However, its occurrence is deemed

relatively less frequent without the influence of environmental effects, grain boundary

particles and creep formation [26, 91, 112]. Incidents of purely mechanical fatigue failure

along grain boundaries have also been documented [114-118]. In general cracking at grain

boundaries may be through one of the following ways [116]:

1) At low and medium plastic strain, grain boundary cracking occurs as a consequence

of impingement of PSBs with grain boundaries as illustrated in Figure 2-23a).

40
2) At high plastic strain, when PSB formation is subdued, fatigue cracks are found to be

preferentially formed from grain boundaries or twin boundaries surface steps formed

at the boundary thereby resulting in cracks as seen in Figure 2-23b).

Figure 2-23: a) Nucleation of flaws along grain boundary [117] b) White light interferograms

showing slip steps formation at grain boundary in fatigue Cu. The dark diagonal

lines parallel to the arrow are fiducial markers whose separation is 100 μm [116]

Furthermore, Kim and Laird [116] determined the possible grain boundary conditions

for crack nucleation at high strain amplitude, if i) the grain boundary is high angle boundary,

ii) the high angle grain boundaries at the free surface lies at a large angle (30°-90°) with the

tensile stress axis and iii) an active slip system site is located at the intersection of the

boundary with the specimen surface. It was also noted that the development of grain

boundary steps was governed by the active slip systems in the neighbouring grains.

On the other hand, crack initiation at twin boundaries, where there is mirror lattice

symmetry across the boundary, has also long been recognized [89, 119-121]. In FCC metals,

even when slip activity is suppressed within the grains at low imposed stress amplitude,

persistent slip bands can be activated exclusively in parallel with or at twin boundaries as a

result of local stress concentration. Such existence of local stress concentration at twin

41
boundary was proven by Wang and Margolin [122]. They have calculated the actual stress

distribution near to and at twin boundaries, and found much increased resolved shear stress

on slip systems parallel to twin boundaries regardless whether the applied stress is tensile or

compressive. Since the applied stress axis does not necessarily form at the same symmetrical

angle between the pair of planes, it may give rise to inhomogeneous deformation between the

two sides of the twin boundary. As a result, enabling the twin boundary to slip and yield the

extrusion-like feature on the surface.

Figure 2-24: Relationship of stress axis and the twin plane (a) Stress axis normal to a twin

boundary (b) Stress axis incline to the twin boundary. This indicates that the

stress axis is crystallographically different in each crystal [122]

With the twin boundary‘s unique configuration, twin boundary cracking and slip bands

are reported to have the tendency to form at every other twin boundary [120]. Considering a

stack of twins, when observing across the boundaries transversely, one may see the lamellae

changes back and forth from one set of twin boundary to another as illustrated in Figure 2-25.

Hence, with the alternative orientation in crystallography, the resultant stress in response to

the applied load also changes accordingly between the two sorts of twin boundaries. Until the

resultant stresses in one type of twin boundary have sufficiently high magnitude, PSBs would

42
develop near these coincidental twin-matrix interfaces and eventually yield cracks. Much

detail of similar trends are also documented on the annealing twin formed in polycrystalline

Cu, Ni and austenitic stainless steel by Neumann [121].

Figure 2-25: Stack of twin boundaries. The white arrows indicate twin boundaries [123]

2.6. Fatigue Crack Behaviour under Pure Compression Fatigue Conditions

The knowledge pool on crack behaviour under compression fatigue, on the other hand,

has always been limited due to the general disbelief in crack nucleation and non-crack-

propagation under pure compression fatigue conditions, however, the investigation of crack

behaviours under the application of uniaxial cyclic compressive loads is not completely

unheard of. As early as the 1960s, Hubbard [10] had already studied the crack growth from a

center-notched plate of 7075-T6 aluminum alloy under cyclic compression. Similar to this

investigation, there were subsequent experimentation [6, 11, 124-130] performed to observe

the crack behaviour from notched plates under pure compression fatigue conditions.

These results suggest that as a notched metallic plate experiences a cyclic compressive

load, cracks could nucleate and grow along the plane of the notch, that is, perpendicular to

the far-field applied compressive stress field. The cracks then grow at a gradual decreasing

rate until they do not propagate any more [6, 11, 126, 127, 131]. Among these researches, it

is argued that residual tensile stress is the basis of this crack initiation and growth. Hence,

43
this issue has been the focus of compression fatigue for a long time. As such, fracture

mechanics is employed to estimate the size of residual tensile zone varying with the cyclic

stress, thereby computing the threshold stress range responsible for crack initiation at the

crack tip [11, 126, 127].

Empirically a compact tension (CT) specimen as shown in Figure 2-26 has been the

most typical specimen design in studying compression fatigue crack. According to Suresh

[131], data from this sample geometry were reproducible, and the crack front obtained was

uniform. The results could also be easily compared with prior studies as this geometry has

been extensively used. In addition, the information obtained is not inferior to those from the

double-edge notched specimens or the notched bend samples employed in former

investigations. Experimental measurements from these types of specimens show that crack

advancement under far-field cyclic compression is essentially governed by local residual

tensile stress. During the loading portion of compressive cyclic stress, a plastic zone is

induced ahead of the notch tip at maximum magnitude of compressive load and its size (r) is

roughly equal to that calculated with Eq. 2-14 under plane stress conditions, where Kmax is

the maximum stress intensity factor.

( ) Eq. 2-14 [68]

Following unloading, the reverse flow therefore generates a zone of residual tensile stresses

at the crack wake and hence causing the crack to initiate [6, 125, 127]. In other words,

compressive stresses could not have induced fatigue crack initiation alone; a net tensile

residual stress must be present in order for crack to initiate in compression fatigue [125, 127].

Therefore, the crack growth behaviour is said to be similar to those occurring in far-field

cyclic tension especially the effect of microstructure on crack opening [26].

44
Figure 2-26: Demonstration of compression fatigue crack in CT specimen

The establishment of a residual tensile stress zone size would set the boundaries for

crack advancement. One of the common observations is that the compression fatigue cracks

nucleated at the notch root of CT specimen stopped growing at a distance of several

millimeters from the notch root [10, 11, 26, 126, 127, 131]. The saturation crack arrest

distance is basically determined [10] to be an exhaustion of the residual tensile zone and

hence leaving little damage at the crack tip when it arrests. Furthermore, the size of the

residual tensile zone is affected greatly by the applied stress condition, i.e., plane strain or

plane stress condition. Plane strain condition results in smaller residual tensile stress field and

thus giving smaller crack advancement in compression than in plane stress. As seen on the

fracture surface, at mid-thickness where plastic constraint is larger and favours a plane strain

condition, a smaller crack length is formed. Whereas on the surface, where plane stress

conditions prevail, the crack propagation is much longer due to the larger plastic zone size.

Other factors that may govern the residual tensile stress zone size are the notch angle or the

notch root radius. An increase of notch angle or notch root radius would lead to larger plastic

zone size and hence a longer saturation crack arrest distance [128, 129]. In addition,

Suresh[131], has established a critical range of compressive loads whereby the crack arrest

distance is independent of the cyclic compressive loads above such loads. Below this load

45
range, crack arresting length is still on the order of the estimated size of the residual tensile

stress region corresponding to the far-field compressive loads.

On one hand, the growth rate of cracks formed under far-field cyclic compressive load

was found to decrease steadily with increasing crack length and is certainly much smaller

than in tensile fatigue [11, 127]. It is explained through the effective stress intensity range

using the concept of crack closure. As the crack length increases the duration of crack

opening in the load cycle decreases and ultimately causing the crack to arrest [127]. This was

made more apparent when comparing cases of divergent applied stress amplitudes at a fixed

mean stress in the negative cyclic stress regime. Certain proportionality is seen between the

amplitude of stress and the total crack length. As the amplitude is reduced, the shorter the

final crack length in return [132]. Subsequently, when linking the threshold value required

for crack to initiate from a notch under compressive with that in tensile cyclic loads, it is also

found the threshold stress required is at least four times larger in compressive than in tensile

[11, 126].

In a more recent compressive fatigue examination documented by Zhao and Jiang [130],

an un-notched tensile bar as opposed to the CT specimen was used in their study. They have

analyzed aluminum fatigue cracking behaviour employing the fatigue model developed by

Smith, Watson and Topper, namely SWT fatigue criterion in its extended interpretation made

by Socie [133] as shown in Eq. 2-15:

Fatigue Parameter (FP): Eq. 2-15 [133]

The σmax and in Eq. 2-15 are the maximum normal stress and normal strain amplitude at

the critical plane defined as material plane that contains highest normal strain amplitude [81,

130, 133]. Despite their efforts in perfecting the fatigue criterion to predict crack direction

and fatigue life of aluminum alloys under full cyclic spectrum, ranging from tension-tension

46
and reversal to compression-compression loading, the crack nucleation direction was still not

properly predicted under negative cyclic stress regime. It was noted that in all compression-

compression cycling cases, cracks nucleated perpendicular to the applied load axis when the

predicted cracking plane was  45∘ to the loading axis. Like in most of the compression

fatigue studies, the deformation mechanisms of fatigue crack nucleation under negative loads

in their report [130] were still not given with clear justification.

The above has demonstrated that most of these works were fracture-mechanics based

studies on fatigue crack behaviours, where there was almost no mention of cyclic

deformation or damage mechanisms, such as dislocation plasticity, microcracking, phase

transformation or creep in these works. Only in one study [134], where the cyclic

deformation of commercially pure zinc under cyclic compressive loading condition was

briefly discussed. It was demonstrated that the main damage mechanisms in commercially

pure zinc, an HCP material of which the predominant plastic deformation is due to twinning

rather than slipping, were owing to the combination of local cracks, deformation twins,

secondary twins, slipping as well as kink banding under cyclic compression testing.

Moreover, as shown in Chapter 1.1.1, the present author has attempted [13] to evaluate

the necessary conditions and micromechanisms for crack nucleation under cyclic

compressive conditions using a notched specimen geometry. Although the idea of residual

stress was not disputed, this work provided additional information for crack nucleation under

pure compression fatigue condition such as the critical stress condition required as well as the

microstructural-related crack nucleation micromechanisms. To reiterate the results, not only a

critical von-Mises stress criterion for crack nucleation was established for these notched

specimens, but crack nucleation was also found to relate to severe plastic deformation. As

such these cracks were seen to nucleate at a close approximation of 45°to the applied normal

stress, where maximum shear stress was found.

47
In summary, the behaviour of compression fatigue is still not extensively explored in

comparison to that in tension fatigue. Among the limited number of compression fatigue

studies, it could also be seen that there are more references on crack growth than crack

initiation under cyclic compression. It is also found that in all these studies, notched

specimens were often utilized. The primary conclusions derived from these publications

reveal that the crack growth rate and final saturation length are determined by specimen

geometry, notch geometry, stress state, stress range and microstructure. General cyclic

deformation and detailed crack nucleation micro-mechanisms are still yet to be determined

especially in the case where macroscopic stress concentration is not present. Further analysis

on these aspects under cyclic compression is therefore required.

2.7. Asymmetrical Behaviour – an Overview of Bauschinger Effect

Bauschinger effect is an asymmetrical phenomenon that was first discovered by Johann

Bauschinger in 1883 [135]. According to Orowan‘s summary [136], Bauschinger effect

describes the lack of corresponding behaviour upon repeating the same loading direction. It

was found the prestrain increased the elastic limit of the material when loaded in same

direction. Or it refers to cases when unequal responses between forward and reverse loading

are observed. These include the plastic prestrain in the forward decreases the elastic limit in

the reverse and/or if the prestrain was large enough the yielding in the reverse direction could

be completely annihilated. Furthermore, in spite of the elastic limit being reduced upon

reverse deformation, it may be increased through alternating directions with magnitude never

exceed the original elastic limit. Overall, investigation of the Bauschinger effect is a study of

the deformation reversibility and mechanism.

By definition cyclic loading is a series of forwarding and reverse loading motion.

Consequently to understand the essence of fatigue resistance, it is imperative to study the

mechanism of Bauschinger to rationalize the work hardening phenomenon therein. For

48
example, the occurrence of cyclic creep and cyclic relaxation may be described from the

basis of the Bauschinger effect. [26]. In addition, various forms of dislocation-related

strengthening mechanisms could be identified and separated. To do so, different

terminologies that are used to quantify Bauschinger effect are given first in the next section.

2.7.1. Terminology

To describe and quantify the Bauschinger effect, it is most common to use "reverse

strain", (εR). The term is defined as the strain that is required to reach the equivalent stress in

the reverse loading as that of the maximum peak stress, σp, in the forward loading spectrum.

The determination of (εR) involves reconstruction of the reverse loading along with the

forward loading based on the magnitude of stress and the accumulated strain as shown in

Figure 2-27. Then, by connecting the maximum stress point from the forward loading curve

with the same stress point on the reverse loading, the reverse strain is therefore obtained or

namely the Bauschinger strain (β) is defined. Additionally, the Bauschinger effect is also

evaluated by the magnitude of "permanent softening", which is expressed in terms of stress,

∆σp. The term is attained from the same Figure 2-27. From the maximum stress of the

forward loading curve, a linear segment is drawn with slope that follows the forward

hardening until such line is found to be parallel with the reverse loading curve. The

difference in stress between the extended segment and the reverse loading curve is defined as

the permanent softening. Three more terms could be found on Figure 2-27, σy, σR, and εp.

These three parameters indicate the yield stress on the forwarding curve, reverse yield stress

on the reverse loading curve as well as the prestrain that represent the strain that was

accumulated in the forward loading curve, respectively.

49
Figure 2-27: Illustration of terminology of Bauschinger effects adopted from Abel [137]

2.7.2. Mechanisms of Bauschinger Effect

The mechanistic reasoning for Bauschinger effects are often offered on the basis of the

change that occur in the dislocation structures or internal stresses. Due to the complexity of

Bauschinger effect, numerous models have been developed to explain the phenomenon.

However, amongst all the theories and models, the dislocation theory and composite model

are most commonly used. They are reviewed and described in the subsequent sections.

2.7.2.1 Dislocation Approach

Dislocation theory was originally proposed by Mott [138] and was later developed by

Seeger [139] to describe the Bauschinger effect. The idea of dislocation theory was the

engendering of long range stress through the formation of dislocation pile up at the barriers,

such as at the grain boundaries or at the front of non-shearable second phase particles. They

are illustrated in Figure 2-28 (a) and (b) respectively.

50
Figure 2-28: Schematic diagrams showing (a) the dislocation pile-ups against a grain

boundary and (b) the interaction between a mobile dislocation and the second

phase particle at different loading stage [140, 141]

At these barriers, dislocation forward motions are prohibited. Dislocation clusters are

hence piled up and formed Lomer-Cottrell* locks at the obstacles, which then hold these

dislocations in position even when the load was withdrawn. Consequently, back stress is

created and favours the motion in the reverse direction [138]. This can mathematically

described in Eq. 2-16.

Eq. 2-16

Where indicates for applied stress, is the frictional stress on the gliding plane and is

the back stress that yielded from dislocation pile up. The equation suggests when the

dislocation pile generates a large magnitude of back stress the stress required for reverse flow

stress is significantly lowered. Hence, in essence, this is the origin of Bauschinger effect.

51
*Lomer-Cottrell locks is a special configuration of dislocations which is sessile and immobile, hence is called ―lock‖
Orowan and his group [142-144] have further elaborated Mott‘s dislocation theory on

the Bauschinger effect based on the permanent softening phenomena that they have observed

in metals such as copper, aluminum, brass, nickel and magnesium. This observation was

made from the change in its stress-strain relation‘s characteristic parabolic form. They have

also noted that upon heating the prestrained sample after the unloaded state, it minimizes the

anisotropy of the material in their stress-strain response. If the heating was carried out above

the recrystallization temperature, Bauschinger effect could be completely annihilated. Such

results have led Orowan to believe that there must be other mechanisms involved behind this

permanent softening as such it is associated with the directional driving resistance of the

dislocations.

The explanation that they offered utilizes modelling of the interaction between

dislocations and non-uniformly distributed obstacles with diverse strength, such as the

permeable (soft) and non-permeable (hard) obstacles. It is said that for the non-permeable

obstacles, the mobility of dislocations is inhibited. As a result, Orowan loops are formed and

back stress hardening is promoted. Hence, higher stress is required to move the blocked

dislocations in the same forward direction. Naturally, the tendency for the dislocations to

move under the reverse deformation would be higher. Subsequently, under the reverse plastic

deformation the back stress hardening would be eliminated by reversed plastic deformation

and result in permanent softening. In contrast, in the case of the permeable obstacles the

dislocations could easily sweep through soft areas with relatively low flow stress where there

is less resistance from these obstacles. The dislocation movements would continue until they

are held against a specific row of obstacles that are closely spaced. Such dislocation activities

leave very small numbers of pile ups. Thus, upon stress reversal, these dislocations would

revert back with small force required. If in the absence of dislocation pile ups, flow stress

would then eventually approach levels obtained during prestraining as the dislocations

encounter another set of impermeable obstacles. Since the model assumed for non-uniformly

52
distributed obstacles with varying strength, the difference between in the flow stresses

between forward and reverse loading would be viewed as the consequence of back stress

hardening. This, in short, was the result of dislocation interacting with the hard particles as

opposed to the minimal resistance given by the soft obstacles during the prestraining process.

Upon load reversal, it would thus be easier for the dislocations to move in the reverse

direction until it again encounters another region of high resistance.

They continued to explore the underlying cause for Bauschinger effect and concluded

that the effect was the results of long range back stresses and the uneven distribution of the

dislocations in a plastically deformed state. Unfortunately, such conclusions were not well

supported. The idea of long range back stresses was deemed as contradiction to the work

hardening theories proposed by other researchers [145-148]. The main argument was the

stability of the deformed state against back-flow during unloading. It was until the

modification made by Brown [140] on Orowan‘s model that had popularized Orowan‘s

explanation on the anisotropy of driving force to dislocation movement. It was said that it

accounts for many features of plastic hysteresis loop.

In general to quantify the back stress within the metallic materials, Ibrahim and

Embury‘s expressions [149] for forward and reverse flow stress, and are often utilized.

represents the peak stress that the material is subjected to in the forward direction before

unloading, namely the in Figure 2-27, where is the flow stress in the reverse direction

as shown in Figure 2-27. They are defined in Eq. 2-17 and Eq. 2-18 respectively.

Eq. 2-17

Eq. 2-18

Similar to Eq. 2-16 is the back stress, whereas and are terms for initial yield stress

and the forest hardening effect from the dislocation interactions, correspondingly. Since the

53
back stress may desist or assist the flow stress depending on whether the loading direction is

forward or reverse, the sign for back stress is assigned accordingly. The permanent softening

is therefore the difference between the and as shown in Eq. 2-19. The back stress could

also be expressed in terms of permanent softening as in Eq. 2-20.

Eq. 2-19

Eq. 2-20

2.7.2.2 Composite Model

The internal stress is also sometimes considered responsible for the rise of Bauschinger

effect [150-158]. Frequently, internal stress mechanisms are explained using composite

models, where each constituent within the composite has significantly different properties

from each other such as their mechanical properties. The idea was first initialized by Heyn

[150], who had generalized materials into small volume of elements that have elastic limits

that differed from each other. Subsequently, if the material is subjected to a load, it is natural

to assume there will be an elastic-plastic mismatch between the different elements. As a

result, it leads to the development of internal stresses within the material and hence gives rise

to the Bauschinger effect. The assumption was strongly supported by Masing‘s experiments

on brass under tension-compression [151-153] as well as the torsion tests on iron, brass [156]

and aluminum wires. It was also commented by Schmid and Boas[155], and Orowan[154]

that the Bauschinger effect was due to the action of internal stresses.

Thompson and Worth [157] and Polakowski and Ripling [158] had extended the theory

by paying closer attention to the two neighbouring grains within the material, which have the

same elastic constants but different elastic limits. With sufficiently large strains, the grain

that has lower yield would experiences plastic deformation when the other grain is still

54
within the elastic deformation. As the applied load is reduced to zero, the two grains would

have deformation incompatibility due to the fact that one grain is in the tension state while

the other is in compression; thereby creating internal stresses near the grain boundary. If the

subsequent load was reversed the internal stresses in the plastically deformed grain would

assist the applied stress and hence reach the yield strength more easily. Consequently the

behaviour is asymmetrical and is said that ―system is softer for the reverse stressing than it

was in the virgin state‖ [137, 157].

Asaro [159] had modeled the two-neighbouring-grain idea by utilizing two elastic-

perfect plastic elements of identical geometry with one element having lower yield strength

than the other. The results indeed followed that of the two-neighbouring-grain‘s approach.

Namely, the soft element will yield before the hard phase provided that if the prestrain is

large enough but smaller than the hard element. When the load is removed different residual

stresses are developed in each element. Continued loading in the reverse direction would

therefore result in lower yield than the forward direction. This phenomenon can be

summarized as the stress-strain response in Figure 2-29 from Wang and Margolin [160],

where W is the stress-strain curve of the weak element that has lower yield strength and

conversely the S represents the curve for stronger elements with higher yield strength.

Assuming both weak and strong component have the same modulus of elasticity, the ―C‖

signifies the stress-strain response of the two elements combination, i.e., the composite. As

such, it is seen that during prestraining when the weak element yields, the bulk materials also

exhibits yielding despite the strong constituents still remains in elastic. After the load is

returned to zero from prestraining, namely, the bulk curves return to zero load (point O), the

weak constituent shows to have the compressive residual stress level of ̅̅̅̅ while the strong

element is under tension residual stress with the magnitude of ̅̅̅̅ . Consequently, the

compressive residual stress within the weak component will assist the deformation upon the

55
opposite loading direction and, as a result, the sooner the composite or the bulk component

will reach to the yield.

In spite of Asaro‘s model being based on the two identically sized elements, it has been

proved to be well applicable to materials that have multiphase structures or the composite

materials such as the results from Wang and Margolis [160].

However, for materials that are of pure phase and have highly symmetrical crystal

structures and containing numerous equivalent slip systems, the internal stress mechanisms

has not found to be predominant, for instance, FCC materials [141]. In these materials, the

deformation is more uniform throughout the entire bulk as each grain experiences almost

equivalent plastic strain during loading. Hence, after the load is removed, the material only

experiences little residual stress. In contrast, for BCC or HCP materials, due to the nature of

these crystal structures, the plastic deformation is not uniform between grains and hence

leaving residual stresses to develop. This in turn will contribute significantly to the

Bauschinger effect.

Figure 2-29: Schematic of stress strain

response of the composite

model [160]

56
3. Methodology and Experimental Procedures

3.1. Experimental Strategy

The following experimental methodology was employed to achieve the objectives of

this study.

(i) Selection of a material and detailed characterization of the general properties of the

selected material, which includes metallography and monotonic tensile and compression

tests.

(ii) Subject the material to different cyclic loading conditions, and analyze various

aspects of cyclic deformation mechanical response and behaviour.

(iii) Documentation of cyclic deformation development by semi in-situ tracing of

surface features at the designated cycle number using 3 different microscopies, namely,

optical microscopy (OM), scanning electron microscopy (SEM) and atomic force

microscopy (AFM) to determine the on-set of crack nucleation.

(iv) Lastly, examine the interior dislocation structure at the termination of the cyclic

tests, using transmission electron microscopy (TEM).

3.2. Material Selection

Table 3-1: List of Impurities in OFHC Cu 101 in ppm [161]

Sb As Bi Cd Fe Pb Mn Ni O P Se Ag S Te Sn Zn
4 5 1 1 10 5 0.5 10 5 3 3 25 15 2 2 1

OFHC Cu is often regarded as the prototype material for fundamental fatigue and cyclic

deformation studies due to the large understanding and knowledge pool on cyclic

deformation behaviour and crack nucleation mechanisms, OFHC Cu 101 in fully annealed

condition is selected to be the material of interest in this project. Such copper is considered as

57
the highest purity of Cu with 99.99wt% of Cu, which according to the ASTM standard has

traces of impurities as listed in Table 3-1.

3.3. Sample Design and Test Set-up

Two types of sample design were utilized in this investigation, Design 1 and Design 2.

The purpose and the actual design of both will be described in detail in this section.

Design 1 was developed to achieve accurate alignment in the monotonic and cyclic tests.

This design takes form of a dumbbell shaped specimen similar to Wang and Laird‘s [162] as

shown in Figure 3-1 with slight modifications. The exact sample design and dimension is

illustrated in Figure 3-2. Correspondingly, a specialized set of grips was devised for

mounting this type of samples on the machine. These grips utilized the same technology as

the Self-alignment Grip for Mechanical Testing with the actual schematic provided in Figure

3-3 [163]. The top and bottom grips are essentially 2 alloy pots, which are to be filled with

Wood‘s Metal. The chemical composition of Wood‘s Metal is listed in Table 3-2. Since the

alloy has low melting temperature of 70°C [164] , it makes sample loading relatively simple

and easy to adjust for alignment. After mounting the sample into one grip with precise

alignment, the assembly is then attached to the top load frame of the testing machine.

Subsequently, it is lowered into the other pot, i.e., the bottom grip, filled with the molten

Wood‘s Metal. The entire setup is then cooled until the alloy solidifies in the bottom grip. In

addition, Wood‘s Metal is also known to have a good surface wettability to copper. As

annealed copper is very soft, such grip devices provide a better gripping than the

conventional hydraulic grips to copper with minimal misalignment. Note that to remove the

sample from the machine for surface observations, a reverse of the loading procedure

described was performed.

58
Figure 3-1: Dumbbell shaped specimen [162]

Figure 3-2: Schematic of sample Design 1

Figure 3-3: Self-alignment grip for mechanical testing [163]

Table 3-2: Chemical composition of Wood‘s Metal in wt% [164]

Bi Pb Sn Cd
50 26.7 13.3 10

59
Design 2 was machined to allow for easy surface observation, characterization and

identification of crack nucleation in the cyclic tests. This type of design has simple

rectangular block geometry with the dimension of 15 x 7 x 7 mm as shown in Figure 3-4.

Such sample design involves using two parallel platens as top and bottom grips and enables

for a simple sample loading and unloading for intermediate surface characterizations.

Furthermore, as the cyclic tests were carried out in compression, not only control of the

alignment, but also specifications of the slenderness ratio were essential to avoid buckling.

Therefore, both of the sample designs, Design 1 and Design 2, also followed ASM handbook

[165] with an aspect ratio of 2.1.

Figure 3-4: Schematic of sample Design 2

In addition, the entire collection of samples was annealed at 625 °C in argon atmosphere

for 3 hrs before all testing. This was done to relieve any residual stress there may be after

sample machining. The machine used for monotonic mechanical tests and cyclic tests was an

MTS frame with 100kN and 50kN load cells respectively. The test parameters were

programmed and monitored by INSTRON 8800 electronic control, computer control panel

with a data acquisition system (FastTrack8800) (refer to Figure 3-5).

60
Figure 3-5: MTS electro-servo hydraulic system

3.4. Microstructure Characterization

Microstructure analysis was achieved by optical microscopy to characterize

microstrctures. The specimens were first cut from the selected tensile bars of Design 1 in the

grip portion. Then the samples were ground successively using sand papers with grit sizes of

400, 600, 800, 1200, and 1200 fine and polished with 1.0 micron diamond suspension

solution. They were etched with an etchant that was composed of 30 mL HCl, 10 mL FeCl3,

120 mL H2O for 15 seconds [166]. The optical microscope used was the Olympus PME3,

which was also equipped with a Nikon Coolpix E995 digital camera for imaging.

3.5. Mechanical Properties Characterization

Tensile and compression testing was carried out to characterize the specimen‘s

mechanical properties. In detail, tensile tests were conducted to obtain the complete tensile

profile of the material while compression tests were performed to evaluate only the

material‘s compressive yield strengths and elastic moduli. This not only ascertained the

determination of fatigue parameters but also to authenticate material‘s isotropic properties.

Both tensile and compression tests were completed using 12.7 mm gage length extensometer

and strain rate of 0.013 mm/sec at room temperature. The samples of Design 1 were tested 3

61
times using 3 separate specimens to obtain an average of each mechanical property for either

compression tests or tensile tests. The data collected were then charted into

tensile/compression profile from which general mechanical properties such as E, σ y, UTS,

and EL% were determined.

3.6. Cyclic Loading Tests

Before commencing fatigue testing, the gauge surface of each sample was ground,

polished and then electropolished. This is done not only to remove the machine marks but

also to facilitate easy surface observation. The electropolishing procedure will be described

in depth in the next section.

Individual samples were then mounted on the MTS frame to carry out stress-controlled

cycling with sinusoidal waveform. Each specimen was loaded with a weighted stress of the

yield strengths pre-determined. These experimental conditions are specified in Table 3-3. It

should be noted that the sample designations reflects the weighted peak stress loads and the

fatigue conditions. For example, a sample designation of 120CF indicates the fatigue

condition is compression fatigue (CF) with the controlling peak compression stress equal to

120% of the yield stress. By the same token, 120TCF would represent tension-compression

fatigue (TCF) condition, i.e., symmetrical fatigue conditions with 120 denotes the

| |= 120% of the yield stress. In addition for symmetrical fatigue conditions, in

order to observe the asymmetrical behaviour, the initial half cycle may start in tension or

compression. The designation is hence either TCF or CTF depending on the sign of the

starting half cycle. Therefore, for all cyclic tests, these sample designations are also referred

as the fatigue conditions hereafter. The selection of R ratio for compression fatigue was

chosen on the basis of smallest σmax but to avoid σmin = 0 so as to have a cyclic range that was

in pure compression only. However, due to the definition of R= , it becomes a reverse of

what the R value generally would be in tensile fatigue. Thus, the R ratios in all cyclic

62
compression testing conditions have a fixed large value of 20. For symmetrical cyclic tests,

the R ratio is thus -1. Loading frequency of 0.5 Hz was chosen for both symmetrical and

compression fatigue.

Table 3-3: Summary of fatigue testing conditions

Pure Compression Test R = 20 Symmetrical Fatigue Test R = -1


Sample Peak *σmin *σmax Stress Sample Peak *σmin *σmax Stress
Designation Stress Range Designation Stress Range
Cond.  Cond. 

80CF 80% σy -23.8 -1.2 22.6 90TCF 90% σy -26.8 26.8 53.6
100CF 100% σy -29.7 -1.5 28.2 90CTF 90% σy -26.8 26.8 53.6
120CF 120% σy -35.6 -1.8 33.8 100TCF 100% σy -29.7 29.7 59.4
140CF 140% σy -41.6 -2.1 39.5 100 CTF 100% σy -29.7 29.7 59.4
160CF 160% σy -47.5 -2.4 45.1 120TCF 120% σy -35.6 35.6 71.2
250CF 250% σy -73.8 -3.7 70.1 120CTF 120% σy -35.6 35.6 71.2
332CF 332% σy -98.6 -4.9 93.7
* All σ are in MPa

The compression cyclic tests were divided into two parts based on the purpose of the

experiments. In the first part of cyclic tests, the cyclic stress-strain responses were obtained

using clip-on exteometers up to 10,000 cycles for conditions of 80CF to 250CF as well as

100TCF, 120TCF. All of these tests were executed with samples of Design 1 configuration.

In the second part, the cyclic deformation development was investigated by semi in-situ

tracing method using various microscopy techniques. Both sample designs were employed in

this part of cyclic tests. During the progression of fatigue tests, samples were removed from

the machine intermittently for examinations and documentation of the surface topography

using both optical microscope (OM) and Scanning Electron Microscope (SEM). For selected

conditions, Electron Backscatter Diffraction (EBSD) scanning in SEM and Atomic Force

63
Microscopy (AFM) measurements were also carried out. For some fatigue conditions, the

cumulative plastic strain were also monitored and recorded using the change in micro-

hardness indent spacing. The technique of measuring the change in micro-hardness indent

spacing will be described in length in the later section. The outline of the tested conditions,

termination of cycle numbers and different characterization methods for this part of cyclic

tests are summarized in Table 3-4. Lastly, an additional cyclic test of the condition 332CF

was carried out up to 100,000 cycles in one test. The sample was then characterized by OM,

SEM, EBSD and micro-hardness indent spacing measurements. The purpose of this

additional test was to further verify some of the surface phenomona that were observed in the

previous surface observations.

64
Table 3-4: Summary of semi in-situ trace conditions

Sample Sample Cycles No. at Surface Characterization Methods


Designation Design Termination
100CF Design 2 100,000 OM, SEM, AFM
Micro-hardness Indent Spacing
120CF Design 1 100,000 Measurements

120CF Design 1 200,000 OM, SEM

140CF Design 1 100,000 SEM


OM, SEM, Micro-hardness Indent
160CF Design 1 100,000 Spacing Measurement
OM, SEM, EBSD, Micro-hardness Indent
250CF Design 2 1,000,000 Spacing Measurement
Micro-hardness Indent Spacing
80CF Design 1 1 Measurement
Micro-hardness Indent Spacing
100CF Design 1 1 Measurement
Micro-hardness Indent Spacing
120CF Design 1 1 Measurement
Micro-hardness Indent Spacing
140CF Design 1 1 Measurement
Micro-hardness Indent Spacing
160CF Design 1 1 Measurement

3.7. Semi-in-situ Observation and Characterization of Surface Evolution Due to Cyclic

Loading

3.7.1. Sample Surface Preparation

To prepare the sample for surface observations, conventional electropolishing was

performed on the samples after they had gone through successive grinding and polishing

process. The electrolyte used for electropolishing is a 10M of H3PO4 [167]. Due to the

irregular shape of the sample grips for Design 1, electroplating tape (3M 470 Electroplating

Tape) was utilized to mask the grip sections before exposed these samples to the electrolyte.

This ensured the polishing process were restricted to only the gauge section surface, but not

65
on the irregular grip sections to avoid complicating the electro-chemical polishing process.

For sample Design 2, the entire sample was submerged without further covering. For both

sample designs, electropolishing was performed with an operating voltage of 1.5 V at room

temperature. The duration of the process was also controlled for about 4 minutes. After

electropolishing, not only the machine marks were removed and hence minimized the risk of

stress concentration at these sites, but also the microstructures were revealed moderately.

This further facilitated an easy idenfication of the cyclic deformation mechanisms in relation

to the microstructures.

3.7.2. Micro-hardness Indent Spacing Measurements

As cyclic creep was expected and to avoid damaging the surface of interest from the

clip-on extensometer, another method was employed to evaluate the plastic strain

accommodation. This was achieved by introducing three parallel lines of microhardness

indentations on the surface of interest before the test as illustrate in Figure 3-6. Each line

contains 7 indents with lo = 2.000 0.001 mm spacing between each other. At each test

interruption, the new spacing between indents were measured and recorded as ln. The result

plastic strain can be converted by the simple equation of . It should be noted that these

microhardness indents not only provide the means to measure the plastic strain

accommodation but also act as the reference point for surface tracing.

66
Figure 3-6: Illustration of microhardness indent spacing measurements

3.7.3. Optical and Scanning Electron Microscopy (including Electron Backscattered

Imaging)

In order to observe the surface topography changes, for most samples the test was

interrupted periodically. The sample was then removed from MTS and examined by optical

and scanning electron microscope (SEM) using Olympus PME3 optical microscope and

Hitachi SU3500 at 5kV respectively. This monitoring was performed carefully through the

whole surface to document the surface morphology evolution before the sample was tested

again. At chosen sites, electron back scattered diffraction (EBSD) mode was also used in the

same SEM with step size of 1μm to examine the local mis-orientation and grain boundary

characteristics.

3.7.4. Atomic Force Microscopy Measurement

Atomic force microscopy (AFM) measurements were performed on 100CF sample to

obtain the topographic images and to measure the surface features at each surface

67
observations. The scans were made using contact mode AFM with silicon nitride tip that has

a radius of curvature of 20nm. With the help of the integrated software, Nanoscope IIIa

controller, the selected sites were traced and profiled at different number of cycles.

3.8. Characterization of Dislocation Evolution Due to Cyclic Loading by Transmission

Electron Microscopy

Traditionally the assessment of microstructure evolution during fatigue is by TEM

observation. Often, the TEM specimens are prepared from the gauge length of the fatigued

sample as demonstrate in Figure 3-7. Sections were cut from the gauge portion or middle

portion perpendicular to the loading axis of Design 1 or Design 2, respectively. Subsequently,

TEM thin foils were made from these sections through further polishing to a thickness of

1μm. Electropolishing was then performed using a Tenupol-5 Twin Jet electropolisher at

settings of single flow and the flow rate of 16 and with an operating voltage of 16.5 V at a

temperature of -15 oC. The process was conducted in an aqueous electrolyte that was made of

30% orthophosphoric acid and 70% distilled water. Afterwards, the samples were examined

using Philips Tecnai F20 operating at 200kV. A comparative approach is adopted for the

TEM analysis. By comparing the TEM bright field images of the fatigued sample and the

freshly annealed ones, the structures produced from the fatigue process and those from the

initial stage are distinguished.

Figure 3-7: TEM specimens prepared from (a) Design 1 & (b) Design2

68
4. Results and Discussion I:
Cyclic Stress-Strain Response and Microstructure Evolution of
Polycrystalline Cu under Pure Compressive Cyclic Loading Condition

4.1. Overview

It is well established that polycrystalline Cu exhibits cyclic hardening and saturation

behaviour under symmetrical fatigue conditions. The cyclic stress-strain curve of

polycrystalline Cu also appears to have three distinctive regions. Having a quasi-plateau in its

CSSC of polycrystalline Cu similar to the true plateau for monocrystalline Cu under

symmetrical fatigue condition indicates that the behaviour of polycrystalline Cu is strongly

dependent on PSBs formations. It was further determined that by imposing a positive mean

stress on the loading spectrum, such PSB formation is suppressed and instead the cell

structure prevails.

From the detailed reviews as summarized in Chapter 2, it is evident that the

investigations on the effect of pure compressive load spectrum on materials‘ cyclic

deformation response are not comprehensive even from the aspects of mean stress effect. As

mentioned in Chapter 1, there is a need in uncovering the underlying micro-mechanisms on

crack nucleation under pure compressive fatigue condition especially from the material‘s

perspective. This should include not only the study of cyclic deformation responses leading

up to the initiation of fatigue cracks but also the comprehensive establishment of cyclic

deformation responses under pure compression fatigue.

The focus of Chapter 4, is therefore to present and discuss our results on the cyclic

deformation response and the microstructure evolution of polycrystalline OFHC Cu samples

under pure compression fatigue condition. In addition, a comparison will be made to reveal

the difference between the above cyclic deformation responses and micro-mechanisms with

those of symmetrical fatigue and tension fatigue. Chapter 5, on the other hand, deals with

topics of crack nucleation under pure compression fatigue.

69
4.2. General Properties of Cu

As any fundamental cyclic deformation studies, general properties of the chosen

material are firstly established prior to the cyclic tests. The microstructural characterization

and mechanical properties of fully OFHC Cu 101 are therefore briefly presented in this

section before discussing the cyclic deformation response of polycrystalline OFHC Cu under

pure compression fatigue condition in greater length.

4.2.1. Microstructure Characterization

Microstructures of fully annealed OFHC Cu 101 are identified by optical microscopy as

shown in Figure 4-1. The microstructure exhibits typical fully annealed copper structure,

namely, equi-axed grain structure with presence of annealing twin boundaries. The average

grain size is determined to be 60μm 5μm.

Figure 4-1: Microstructure of fully annealed OFHC Cu 101

4.2.2. Mechanical Properties

Mechanical testing was carried out to determine the mechanical properties of fully

annealed polycrystalline OFHC Cu. The results are shown as follows:

70
Figure 4-2: Tensile and compression profiles for fully annealed polycrystalline OFHC Cu 101

71
Figure 4-2 summarizes stress-strain curves that are obtained from tensile and compression tests.

They are shown as FA_Ten and FA_Comp, respectively, with subsequent number representing

the test number. For tensile tests the profiles were obtained up to complete fracture whereas

compression tests were terminated up to 4% engineering strain. Based on these profiles in Figure

4-2, the tensile properties of fully annealed polycrystalline OFHC Cu are summarized in Table

4-1. The yield strength and elastic modulus obtained from the compression test are also

presented in the comparative manner with these obtained from the tensile test in Table 4-2.

Although the overall profiles between tensile and compression tests seem relatively consistent,

there are still minor discrepancies noted between the two, such as the difference in the yield

strength between the two tests. These will be discussed in detail in Chapter 6.

Table 4-1: Tensile Properties

Elongation
UTS to Fracture
Samples Y. S[MPa] E [GPa]
[MPa]
[%]
Quench-Tempered 30.6 115 214 74

Table 4-2: Tensile vs. Compression Test Results

Tensile Compression

Yield Elastic Yield Elastic


Strength Modulus Strength Modulus
[MPa] [GPa] [MPa] [GPa]
30.6 115 -28.8 115

4.3. Cyclic Response

To study the cyclic response, fatigue conditions of sample 80CF to 250CF as listed in Table

3-3 were carried out. For comparison, two symmetrical fatigue conditions of 100TCF and

72
120TCF were also conducted. Systematic analyses of the mechanical stress-strain responses and

documentation of the microstructural evolutions are presented first. Discussions on the

comparison of the overall cyclic compression responses to that of the symmetrical fatigue as well

as correlation of the microstructure evolution with the CSS response are given in the subsequent

sections.

4.3.1. Cyclic Stress Strain Response

From 6 tested pure compression fatigue conditions, 4 conditions were selected to represent

the overall hysteresis loop behaviours, 100CF, 120CF, 140CF, and 160CF. Summary of these

hysteresis loops under different conditions are shown in Figure 4-3 (a) – (d). Two interesting

phenomena are observed in Figure 4-3 (a) – (d): (1) There is a tremendous amount of uniaxial

plastic deformation, , accommodated in the 1st cycle for loading conditions above 80CF.

As such, the magnitude of this plastic strain in the first half cycle increases with the increase in

applied peak stress. However, in the subsequent cycles, it is observed that the hysteresis loop

becomes almost a linear elastic response, whereby almost no appreciable loop width is detected

from cycle No. 5 until the termination of the fatigue tests, i.e., 10,000 cycles. Using the

hysteresis loops obtained for 160CF as examples, Figure 4-4 (a) – (d) demonstrate such

evolution from cycle 1 to cycle 10; and (2) a clear mean strain shift was noticed as indicated on

each of the figures in Figure 4-3. These observations, however, were not noted in the

symmetrical fatigue.

73
Figure 4-3: Summary of hysteresis loop development in compression fatigue for peak stress conditions of (a) 100CF, (b) 120CF,
(c) 140CF and (d) 160CF

74
Figure 4-4: Illustration of evolution for 160CF at (a) cycle 1 (b) cycle 2 (c) cycle 5 (d) cycle 10

75
Figure 4-5: Comparison of hysteresis loop between symmetrical fatigue and compression fatigue for peak stress conditions of 120
% at different cycles. (a) cycle 1 (b) cycle 5 (c) cycle 10 and (d) cycle 10,000

76
Figure 4-6: Illustration of evolution for 160CF at (a) cycle 1 (b) cycle 2 (c) cycle 5 (d) cycle 10 and (e) – (h) the respective
magnified selected area

77
When the individual hysteresis loops are selected from both fatigue conditions, i. e. both

TCF and CF tests under the same weighted peak stress of 120% , and plotted against the

tensile strain axis in the same graph, the above mentioned dissimilarities between the two

conditions are clearly revealed. Figure 4-5 (a) shows the hysteresis loops of 1st cycle for both

conditions. For this 120TCF condition, it is seen that the material has accommodated two

types of plastic strain, unidirectional plastic strain ( ) and cyclic plastic strain ( ), see

Figure 4-5 (a). In contrast, the material accommodates almost zero ( ) but a large,

in the 1st cycle for 120CF condition. As the cycle progresses, Figure 4-5 (b) and (c), there is

only accommodation of in TCF conditions, whereas the same cyclic plastic strain

accommodation was not noted in the CF conditions. At the 10,000th cycles, both conditions

have shown that each has obtained almost zero plastic strain. Furthermore, from Figure 4-5 (a)

to (d), it can be seen that there is no mean strain shifting for TCF condition, whereas the shift

in mean strain is clearly noted for CF condition.

To reveal additional information such as the ―transient cyclic behaviour‖ (i.e., cyclic

hardening/softening curve) and the aforementioned cyclic creep phenomenon, further

analyses were performed on these hysteresis loops. These are summarized in the succeeding

sections.

4.3.2. Cyclic Hardening Curve and Cyclic Stress Strain Curve

Although the above analysis did not show the obvious cyclic plastic strain ( or loop

width) for CF conditions, detailed analyses on these hysteresis loops have shown that a very

small magnitude of plastic strain is indeed accommodated in each and every cycle as

illustrated in Figure 4-6. Figure 4-6 (a) – (d) present the hysteresis loops of cycle 1, cycle 2,

cycle 5 and cycle 10 of 160CF condition respectively. Figure 4-6 (e) – (h) shows the

magnified selected region in Figure 4-6 (a) – (d). The observed has revealed that in the first

two cycles the material accommodates not only but also . Both strains decrease as

78
the number of cycle increases. Until cycle no. 5, where the hysteresis loop behaviour was

observed to be almost a close loop with very small magnitudes of , and .

To monitor the cyclic behaviour, the per cycle, is plotted against the cycle number.

Using 3 selected pure compression fatigue conditions as examples, Figure 4-7 shows εpl

decreases with the increase in number of cycles and thus revealing cyclic hardening

behaviour. These cyclic hardening curves are further shown comparatively with the cyclic

hardening curve for the symmetrical fatigue conditions in Figure 4-7 (a) & (b). Unlike the

TCF condition in Figure 4-7 (b), the CF samples in Figure 4-7 (a) are shown to accommodate

a plastic strain much smaller than that by the TCF samples, not only in the first cycle but also

throughout the entire 10,000 cycle test. As seen on the ordinate in Figure 4-7 (a), the plastic

strain of the first cycle of the 120CF sample is approximately 1.5x10-5, whereas in Figure 4-7

(b), this value is 1.6x 10-3 for 120TCF, whereas at the 10,000th cycle, the value for 120CF is

1.3 x 10-6 and the value for TCF is 7 x 10-5.

Generally the cyclic hardening behaviour of symmetrical fatigue condition is known to

reach saturation after a certain number of cycles. From Figure 4-7 (b), saturation is indeed

observed for the TCF conditions at around 1000 cycles. The cyclic hardening curve of the CF

condition on the other hand is rather different. From Figure 4-7 (a), one can see clearly that,

although the plastic strain value is very small and still on a descending trend after 10 cycles,

such decrease is almost negligible if compared with that shown for TCF samples in Figure

4-7 (b). Thus saturation of CF samples seems to be reached after approximately 10 cycles,

Figure 4-7 (a). As a specific example, for the 250CF condition, the plastic strain is measured

as 3.9 x 10-5 at the 10th cycle. After 10,000 cycles, for the same sample, the plastic strain is

still at the same order of magnitude with the actual value of 3.0 x 10-5.

79
Figure 4-7: Summary of cyclic hardening response for (a) selected pure compression fatigue
conditions, and (b) symmetrical fatigue conditions

For cyclic stress strain curve (CSSC) constructions, the plastic strain value at the

10,000th cycle from each of CF and TCF conditions is taken as the saturation point. The

CSSC of pure compression fatigue is plotted along with two data points for samples tested

under a symmetrical loading condition. It is observed that the CSSC of CF is almost linear

80
when the stress condition is higher than 120CF as such the plastic strain range increases with

increase of peak stress conditions. It is also found that the CSSC of pure compression fatigue

conditions exhibits no apparent plateau. Furthermore, under the same applied stress

amplitude, the saturated plastic strain range of pure compression fatigue test is noted to be

much smaller than that of the symmetrical conditions. For example, for the 120CF sample,

the saturation plastic strain is approximately 7x10-6, whereas for 120TCF sample the strain

level is 7x10-5. Apparently, a much bigger total strain was accommodated, or accumulated, in

the TCF sample than in the CF sample. This would lead to a different microstructure as will

be revealed in later discussions.

Figure 4-8: CSSC of polycrystalline Cu under pure compression fatigue

4.3.3. Cyclic Creep

In addition to monitoring the plastic strain amplitude, the mean strain shift with respect

to time was also measured. As mentioned in section 4.3.1, a clear mean strain shift was

noticed, Figure 4-9 illustrates such cyclic creep strain evolution. Mean strain is observed to

81
increase rapidly at the initial 2000 sec (1000 cycles) and it reaches a steady state creep status

from 2000 sec and onwards for peak stress from 100CF to 160CF, Figure 4-9. For these peak

stress conditions, it is also found that there is a net creep strain of approximately 0.12% after

10,000 cycles. Figure 4-10 further reflects that the majority of the creep strain comes from

the 1st cycle. This inarguably is the result of the large εpl,uni that was observed in Figure 4-3 (a)

– (d). Upon the removal of the cyclic creep strain from 1st cycle, the total creep strain is

almost constant at the measured net creep strain of 0.12%. To quantify the large one

dimensional strain in the first cycle, additional tests were carried out specifically for one

cycle only to measure the microhardness indentation spacing after 1st cycle. The results are

summarized in Figure 4-11. It was found that the unidirectional plastic strain accommodated

in the first cycle increases linearly with the peak stress condition as illustrated by Figure 4-11.

Figure 4-9: Cyclic creep curves of selected conditions plotted against time

82
Figure 4-10: Total cyclic creep strain against peak stress conditions with and without the 1st
cycle

Figure 4-11: εpl in the 1st cycle

83
4.4. Surface Morphology Evolution

The surfaces of the fatigue specimens of different conditions and at different number of

cycles were observed by optical microscope, SEM and AFM. These were performed to

reveal the microstructural change in relation to the cyclic response. The results are illustrated

in the consecutive sections.

4.4.1. Optical and SEM Observations

The evolution of surface morphology was documented at different number of cycles for

different peak stress conditions. From the extensive optical images, Figure 4-12 to Figure

4-15, it is observed that specimen surfaces of 100CF, 120CF, 160CF and 250CF are all

roughened as the number of cycles increased. This change in morphology was clearly noticed

in as early as the 1st cycle, which are demonstrated by Figure 4-12 and Figure 4-15. The level

of roughness was also indicative by the corresponding εpl, i.e., the higher the εpl the rougher

the sample surface.

To detect the micro-features that give rise to the surface roughening, samples were also

examined by SEM, which reveals that the surface is roughened by features of grain boundary

extrusion as well as slip band formation during early cycling. Such can be demonstrated

using SEM images taken for 160CF as shown in Figure 4-16 (a) and Figure 4-17 (a). As

cycling continued, the grain boundary extrusion appeared to have become more severe,

Figure 4-16 (b) to (c) and Figure 4-17 (b) to (c). On the other hand, both the morphology, the

number and severity of the slip bands remained almost the same as cycle progressed, Figure

4-16 (a) – (c) and Figure 4-17(a) – (c). The phenomena observed here for the 160CF sample

is found to be also true for samples tested under different peak stress conditions. For example,

Figure 4-18 shows the surface morphology for another condition, 250CF, at higher number

of cycles. The observed slip bands show that even at this high peak stress condition and after

a prolong period of cycling, the slip bands do not appear to be pronounced.

84
Figure 4-12: Optical images revealing the surface microstructures evolutions of 100CF sample at different number of cycles at low
magnification. (a) 0 cycle (b) 1 cycle (c) 100,000 cycles

Figure 4-13: Optical images revealing the surface microstructures evolutions of 120CF sample at different number of cycles at low
magnification. (a) 0 cycle (b) 100 cycles (c) 100,000 cycles

85
Figure 4-14: Optical images revealing the surface microstructures evolutions of 160CF sample at different number of cycles at low
magnification. (a) 0 cycle (b) 100 cycles (c) 100,000 cycles

Figure 4-15: Optical images revealing the surface microstructures evolutions of 250CF sample at different number of cycles at low
magnification. (a) 0 cycle (b) 1 cycle (c) 10,000 cycles

86
Figure 4-16: SEM images revealing the surface microstructures evolutions of 160CF sample at different number of cycles at high
magnification for site No. 1. (a) 100 cycles (b) 15,000 cycles (c) 100,000 cycles (Note: the development of the grain
boundary extrusion and the lack of major development of slip bands)

87
Figure 4-17: SEM images revealing the surface microstructures evolutions of 160CF sample at different number of cycles at high
magnification for site No. 2. (a) 5000 cycles (b) 15,000 cycles (c) 100,000 cycles (Note: the grain boundary extrusion)

Figure 4-18: SEM images revealing the surface morphology of 250CF sample at 150,000 cycles
(Note: the right hand is the framed region in (a) at a higher magnification)

88
Moreover, comparing Figure 4-16(a), (b) and (c) or Figure 4-17 (a), (b) and (c), it

appears that, although the grain boundary offset increased with the number of cycles, the slip

band offset did not appear to have increased in any apparent form. To verify further the

aforementioned phenomena and to quantify both the grain boundary offset and slip band

offset, AFM was employed and the results are presented in next section.

4.4.2. AFM Observations

To present our AFM results, first a schematic of the grain boundary offset height

measurement is illustrated in Figure 4-19. As an example, the AFM monitoring was carried

out for the condition of 100CF at different cycle numbers. On the surface of this sample, a

total of 6 spots, or locations, were traced for specific numbers of cycles. Using one of the six

spots, Spot 6, as an example, the grain boundary offset was measured after the sample was

cycled to 5 different numbers of cycles. The results are demonstrated in Figure 4-20, where a

series of offset height profiles along a grain boundary are shown in Figure 4-20 (a) through

(e). The corresponding grain boundary morphologies are shown in Figure 4-20 (f)-(j) at the

corresponding respective number of cycles. These figures reveal the details of the surface

topography evolutions of the offsets due to grain boundary extrusion. In Figure 4-20 (a)-(e),

the profile of offsets along the grain boundary are shown from cycle 1 to cycle 10,000.

Having the ordinate being the offset height measurement under the same scale for all

measurements, it is clear that with the increase in the number of cycles the intensity of the

offset height increases as well. Based on the evolution of the profile, it indeed emphasizes

that the grain boundary extrusion is the major cause for the surface roughening.

89
Figure 4-19: Illustration of grain boundary extrusion

To further summarize these observations, actual numerical values of the grain boundary

offset profiles recorded by AFM at Spots 1 and 2 are plotted in Figure 4-21, where Figure

4-21(a) demonstrates the evolution of grain boundary offset height, and, for comparison,

Figure 4-21(b) shows the results obtained from the offset height measured along a slip band

near Spot 4. From Figure 4-21 (a), it is evident that the grain boundary offset advances with

increase in the number of cycles. For the offset height measured along the slip band, Figure

4-21(b), almost a constant value was measured as the cycle progressed. In other words, there

were no further developments of the slip bands and thus no formation of persistent slip bands

as the number of cycles increased.

90
Figure 4-20: AFM profile results and corresponding images at a grain boundary: (a)-(e): the profile measurements at different
numbers of cycles after cycle number of 1, 100, 1000, 10,000 and 100,000 cycles, respectively; (f-j) the corresponding
images.

91
Figure 4-21: (a) Measurements of grain boundary offset height vs. cycle no. , at AFM
observation Spots 1 and 2, and (b) Slip band offset height measurements vs. cycle
no., at Spot 4
92
4.5. Dislocation Evolution
To examine the dislocation behaviours, TEM investigation was carried out for different

peak stress conditions and at different chosen number of cycles. Figure 4-22(a) shows the

TEM image of the material before testing. As with all the fully annealed copper, only minor

dislocation debris was detected in the as received condition. In Figure 4-22(b), the

dislocation structure of 160CF condition after 1 cycle was examined. This was carried out to

see the dislocation activities since a large unidirectional plastic deformation was measured in

the first cycle from the cyclic deformation stress-strain response. It is seen from Figure

4-22(b) that the cell structures already formed after the first cycle with rather loose

dislocation entanglement as cell walls. These cells are about 2μm in size. It is interesting to

note that even after ten thousand more cycles, the cell structure is still very similar to the one

after the 1st cycle as shown in Figure 4-22(b) & (c). The dislocations still exhibit a loose cell

structures with an approximate cell size of 2μm, Figure 4-22(c), and the overall cell

structures are still not mature, Figure 4-22(c). The cell walls are relatively translucent and do

not compose of dense dislocation structures despite that the sample was tested under such a

high peak stress load and even after the long cycling period.

93
Figure 4-22: Dislocation structure (a) before fatigue testing, (b) after fatigue tested for 1
cycle at 160CF, (c) after fatigue tested at 160CF to 10,000 cycles, (d) after
fatigue tested at 100CF to 10,000 cycles

94
Furthermore, TEM observation was also made for 100CF after 10,000 cycles, and is

presented in Figure 4-22(d). When the three TEM images of different peak stress conditions

were compared, Figure 4-22(a), (c) & (d), it is observed that there is an increase in

dislocation density with the increase of peak stress condition despite the increase may be

moderate. The observed also demonstrate that even at this low stress condition there is still a

tendency for dislocations to form cell structure. As demonstrated in Figure 4-22(d), for the

100CF condition after 10,000 cycles, despite the dislocation density is much smaller than the

160CF after the same number of cycles, these dislocations seem to have already networked

themselves into that of the cell formation.

4.6. Analysis on the Cyclic Stress Strain Response

The cyclic stress-strain response of the fully annealed polycrystalline copper under

symmetrical fatigue is generally found to be cyclic hardening. Such hardening is attributed to

the dislocation activities in the material. It was also found that, depending on the applied

stress or stain amplitude, the cyclic hardening may be gradual or rapid until its eventual

cyclic saturation [16, 75]. In fact, the cyclic hardening behaviour of polycrystalline Cu under

certain peak stress conditions could go on for as long as 104 cycles [16]. This was indeed

observed in the two symmetrically fatigue tested samples in the present investigation. In the

present case, the plastic strain range started at 1.6 x 10-3 and 6 x 10-4 for 120TCF and

100TCF, respectively. The observed cyclic hardening ranges for both samples lasted for

about 103 cycles until both reached saturation. Such gradual hardening behaviour was also

observed on the cyclic hardening curve of single crystal Cu samples under pure tension

fatigue condition, when the applied shear stress amplitude, τa, was below 21MPa [75]. For

shear stress amplitude higher than 21 MPa, it was seen that the samples rapidly hardened in

the first 10 cycles followed by cyclic softening and then an inexpressive cyclic hardening as

shown in Figure 4-23 [75]. By the consideration of Schmid, Sach and Taylor‘s model, the

equivalent normal stress for this transition change in polycrystalline would be 42MPa,

95
47MPa, and 64.3MPa respectively.

Figure 4-23: Cyclic hardening/softening curves of samples tested in pure tension fatigue [75]

The current tested conditions by contrast should fall within the low applied stress

amplitude and hence it should show gradual hardening behaviour. The current results of pure

compression fatigue tests show that there is indeed a cyclic hardening behaviour. However,

as the initial level of plastic strain range, or loop width, was already at much lower level, i.e.,

in the order of 10-5 for all tested conditions, the cyclic hardening behaviour was observed to

take place only within a small number of cycles and to a similar strain range, i.e. still in the

order of 10-5. In fact, it was seen that such a hardening-to-saturation transition occurred in

less than 10 cycles. The saturated plastic strain amplitudes of the pure compression fatigue

samples were also found to be smaller than that of the symmetrical fatigue under the same

stress amplitude, Figure 4-8. This ultimately reflects the effect of compressive mean stress as

such, under the employed pure compression fatigue conditions, the cyclic plastic strain

accommodation is highly suppressed. Hence, as shown by our results, a shift of CSSC to the

left side of the saturated plastic strain spectrum is detected, Figure 4-8, when compared with

96
the CSSC from the symmetrical fatigue testing.

In addition to the cyclic plastic strain, the observed phenomena indicate also that the

material accommodated a relatively significant amount of unidirectional plastic strain, i.e.

cyclic creep strain. This unidirectional strain was comprised of two parts: a large

unidirectional plastic strain obtained in the first cycle and the net shift of the mean strain in

the remaining cycles. The magnitude of the unidirectional plastic strain in first cycle was also

found to increase with increasing applied peak stress as demonstrated in Figure 4-10. These

were apparently the results of mean stress effect as commented by Kunz et al [168]. However,

when comparing the cyclic creep strain rate obtained in our case with that attained from the

pure tension fatigue of single crystalline samples [75], an interesting difference has been

realized. It was found that the cyclic creep strain rate always increases above τa > 19MPa as a

result of the increase in mean stress under pure tension fatigue for single crystal Cu [75]. The

same observation was not seen in this study. In the tested conditions where cyclic creep was

observed, the creep strain rate reaches steady state for peak stress levels that were higher than

100CF, Figure 4-9. Nonetheless, the overall trend shows the main mechanism for plastic

strain accommodation under pure compression fatigue conditions is through cyclic creep

strain instead of cyclic plastic strain.

97
Figure 4-24: Cyclic creep curves of samples tested in pure tension fatigue [75]

Moreover, it is also believed that hysteresis loop evolution of tension-tension fatigue for

polycrystalline copper would have different results than in the current case. As it was noted

by Eckert et al.[58], there exists an asymmetry between tension and compression when

testing polycrystalline copper under symmetrical stress-controlled fatigue. Depending on the

nature of the first stroke, i.e., tension-start or compression-start, the creep behaviour is

different. This invariably implies that the cyclic behaviour between the pulsating tension

fatigue and the pure compression fatigue for polycrystalline copper should be different.

4.7. Correlation of the Microstructure Evolution and the CSS Response

Since the cyclic deformation response of a material is the reflection of the

microstructural evolution within the material upon fatigue, the explanations for

understanding the stress-strain responses are no exceptions. In the current case of pure

compression fatigue, the small saturated plastic strain range obtained in each condition is

reflected by the behaviour of hysteresis loop development which becomes almost elastic once

98
the saturation is reached. Such observation coincides not only with the present TEM results

that show little dislocation activities in the cycled samples but also with the surface

observations that demonstrate the lack of intensive slip behaviour.

As indicated above, all the CSS responses obtained in this study include a large cyclic

creep (uniaxial directional strain) and a small amount of cyclic plastic strain (the hysteresis

loop width). It is interesting to note that, with the accommodation of these strain terms, there

is little dislocation activity once the dislocations have taken the arrangement in the first cycle.

As observed in the optical and TEM micrographs of 160CF condition, Figure 4-22 (b), the

unidirectional strain in the first cycle is simply the result of both grain boundary extrusion on

the surface and dislocation cell structure formation in the interior. With further cycling, the

surface continues to be roughened by grain boundary extrusion, Figure 4-12 (b) & (c) and

Figure 4-14(b) & (c). However, in terms of the dislocation activities, the results show only

that the cell walls have evolved from lose entanglement to more compact and thinner walls.

In addition to the above facts and more significantly, no PSBs were detected both from TEM

investigations and surface observations. As it can be seen from Figure 4-18, even after the

sample was fatigued at much higher peak stress condition and cycled after a prolonged period,

the same persistent slip band morphology, shown in Figure 4-25 as a typical example, was

not noticed. These experimental evidences validate convincingly the following conclusion:

the main mechanism of plastic strain accommodation in the present compression fatigue

condition is grain boundary extrusion.

99
Figure 4-25: Typical persistent slip bands
morphology on the surface [19]

The structure of PSBs generally consists of ladder-like dislocation structures surrounded

by veins. They are commonly found in copper samples, both single and polycrystalline

samples, that are cycled under symmetrical fatigue. In the case of easy-glide single crystal

samples, upon PSB formation, the resulting saturated cyclic plastic strain range falls within a

specific span. However, it was also reported that when the load spectrum was offset from the

symmetrical condition, even with a relatively small positive mean stress in the load spectrum,

such development of PSB is suppressed [75]. In return, only the formation of homogenous

cell structure or veins prevailed [75]. Exact phenomena were observed in the present

compression fatigue study but in polycrystalline samples. Where there is a mean compression

stress, the tendency for cell structure formation is obvious. Specifically, our results show that

such cell structures were already formed in the first cycle and their further development was

rather limited hence yielding in such a low saturation plastic strain.

The grain boundary extrusion, on the other hand, continuously evolved with the increase

in the number of cycles. This is further verified from the AFM trace and TEM observation of

100CF condition after 10,000 cycles. Even after the accommodation of the relatively large

unidirectional strain in the first cycle then the cyclic creep strain, i.e. the net mean strain shift

and the cyclic plastic strain in each cycle at saturated plastic strain range of ( 2.8 x 10-6),

only the phenomenon of grain boundary extrusion and its intensification are still observed.

100
There was almost no evidence of slip behaviour development in both TEM observations and

the surface trace. This again shows implicitly that grain boundary extrusion is the main strain

accommodation mechanism that gives rise to the mean strain shift. It is interesting to point

out that similar phenomena were reported by Kim and Laird [115, 116, 169] under very

different testing conditions with polycrystalline Cu samples, i.e. the very high strain push-

pull fatigue. In that investigation, it is observed that PSBs were suppressed when grain

boundary extrusion prevails. However, they had further noted such grain boundary extrusion

occurs at where active slip trace was observed. As such grain boundary extrusions were seen

to be associated with the slip process. This differs to our observation. The current results

show that there is little dislocation activities in the material through TEM observation and

there is apparent lack of slip behaviours on the surface. Therefore the mechanism of grain

boundary extrusions that is observed in the present investigation appears different from that

in Kim and Laird studies [115, 116, 169] and warrants further investigation. In addition,

since the formation of PSBs requires intensive dislocation movements, the lack of intensive

slip behaviours and little dislocation activities in the current results would suggest that the

formation of PSBs is less likely.

Nonetheless, the observation of grain boundary extrusion and the result of cyclic creep

have led to the comparison between the conventional static creep and the present cyclic creep

case. When polycrystalline copper is loaded under the static load conditions with the stress

level comparable to the present fatigue peak stresses, through the use of deformation map,

the dominant creep mechanism is found to be Coble creep[170]. This type of creep is often

introduced through grain boundary sliding and may eventually cause void formation along

grain boundaries. It is a thermally activated phenomenon that is attributed to the diffusion of

atoms along grain boundaries. Unlike the Coble creep, the current observed cyclic creep is

proven to be the result of grain boundary extrusion.

101
4.8. Summary

In this chapter, cyclic deformation responses of OHFC polycrystalline Cu samples under

pure compression fatigue have been presented and characterized. The intention of this

chapter is to reveal the differences between the compression fatigue condition and these in

symmetrical fatigue or tension fatigue conditions. The major findings and observations from

this investigation are summarized as follows.

(i) Overall, the response and microstructural evolutions of OFHC copper under

pure compression cycling condition exhibits rather dissimilar behaviour

compared to those under general push-pull fatigue.

(ii) Under pure compression fatigue condition, cyclic creep was observed to be the

major form of plastic strain accommodation. There was no major cyclic plastic

strain accommodation detected, even if when the applied stress levels were

high during such a fatigue condition.

(iii) There was little dislocation activities and lack of slip behaviour detected under

cyclic compression.

(iv) The cyclic creep was observed for peak stress condition above 80CF, and it is

associated with the grain boundary extrusion on the surface instead of the

typical dislocation-slip activities. As such grain boundary extrusions advances

as the number of cycle increases.

(v) Although cyclic saturation was observed, the saturated cyclic plastic strain

amplitude is found to be much lower than that of the symmetrical fatigue.

There is no plateau of the CSSC under pure compression fatigue. Most

interestingly, no PSBs were detected under the current testing conditions.

(vi) The testing material was found to be hardened within 10 cycles when cycled

under pure compression loads, unlike its behaviour in push-pull fatigue with the

same stress amplitude.

102
(vii) Cyclic compression deformation results in dislocation cell structures, which

take their form mainly in the first cycle. Thus, under the testing condition, the

maturity of the cell structure is primarily governed by the peak stress load.

Overall, OFHC copper shows to have a rather dissimilar response and microstructural

evolutions under pure compression cycling condition than its behaviour under the general

push-pull fatigue. Our investigation on whether or not compression fatigue would eventually

lead to fatigue crack formation and if ―yes‖ what is the mechanism will be presented in the

following chapter.

103
5. Results and Discussion II:
Investigation of Crack Nucleation Mechanisms of Polycrystalline Cu
under Pure Compressive Cyclic Loading Condition

5.1. Overview

It is well known that fatigue crack nucleation mechanisms in general is associated with

certain form of stress concentration sites such may be surface roughness, defects or notches.

In the case where these surface irregularities are non-existent, namely, nominally defect-free

pure metals, the stress concentration site may arrive from the formation of persistent slip

bands (PSBs) - the most unique feature of cyclic plasticity resulting from cycling, in both

single crystals [171-175] and polycrystals [19, 106, 117, 176]. As such the cracks are

initiated from the root of the intrusions of the PSBs and/or from the impingement of PSBs

with grain boundaries [117, 177, 178].

It was also seen that cyclic deformation may not always lead to formation of persistent

slip bands. Depending on the loading conditions such as the degree of asymmetry of the

loading spectrum or the stress amplitude or the slip character of the material or even the

testing environment, PSBs may or may not emerge from cyclic loading [75, 115, 116]. For

instance, in the presence of mean stress, Lukas and Kunz [75] have determined that the

formation of PSBs only exists for small mean stress conditions for monocrystalline Cu

oriented in easy-glide direction as shown in Figure 2-13. They have further demonstrated that

in the case where PSBs were absent under the influence of positive mean stress, cell

structures were the dominant dislocation morphology. As a result, fatigue microcracks were

found to be linked with the sliding of the slabs of cells in single crystal Cu.

On the other hand, for polycrystalline materials, when PSB formation is subdued,

fatigue cracks may be preferentially formed from grain boundaries or twin boundaries. For

example, the grain boundary cracking in Kim and Laird‘s investigation [115, 116] of crack

nucleation in high strain fatigue under symmetrical loading condition. As demonstrated in

104
Chapter 2, they have established the three necessary grain boundary conditions for crack

nucleation, i.e., i) the grain boundaries are high angle boundaries, ii) the high angle grain

boundaries at the free surface lie at large angles (30°-90°) with the loading axis and iii) an

active slip system site is located at the intersection of the boundary with the specimen surface.

As presented in Chapter 1.1.1, it was observed from our previous investigation on

notched specimens [13] that crack nucleation is possible under pure compression fatigue

provided a stress concentration is present. It was further found that fatigue crack nucleates

when a certain level of von Mises stress is reached at the notch root. Due to the presence of a

notch, a 3D stress state is developed at the notch root. This gives rise of a localized von

Mises Stress that may inflict a severe localized plastic deformation at the notch front. Thus,

the crack initiation process is observed to be associated with plastic deformation including

grain rotation. As a result, the crack was observed to nucleate close to an approximate 45°

angle with the loading axis where the maximum shear stress is found. In Chapter 4, on the

other hand, it is demonstrated that when a smooth bar was involved, the main cyclic

deformation mechanism under pure compression fatigue has always been cyclic creep, which

resulted in grain boundary extrusions instead of PSB formation [179]. However, whether

such grain boundary activities would lead to the eventual crack nucleation warrants further

investigation. The focus of the present chapter is therefore to continue to observe such

phenomena at longer cycles as well as at higher stress conditions, to verify the possibility of

crack nucleation and to determine the crack initiation mechanism under pure compression

fatigue if cracks do form.

5.2. Cyclic Creep

Before examining the microcrack nucleation under pure compression fatigue conditions,

the cyclic deformation responses of polycrystalline Cu are first revisited. As presented in our

Chapter 4 and in Ref. [179], cyclic creep was observed under pure compression fatigue.

These cyclic creep curves were established using the first type of sample design. An

105
extensometer was utilized to measure the strain up to 10,000 cycles for these samples for

conditions of 100CF to 160CF. Figure 5-1 gives the cyclic creep curves for all these samples,

which were shown in Chapter 4 and in Ref. [179]. Along these cyclic creep curves in Figure

5-1, two additional conditions, 160CF and 250CF are also presented. For these two

conditions, the creep measurement was achieved by employing the micro-indentation method

on the block samples up to 100,000 cycles. To distinguish the two 160CF cyclic creep curve

results (one measured with extensometer and the other with indentation marks), 160CF_No1

represents the former whereas 160CF_No2 denotes later.

Figure 5-1: Cyclic creep curves vs. number of cycles

In Figure 5-1, cyclic mean strain is plotted against the number of cycles. It is observed

that much like the previous conditions of 100CF to 160CF reported in both Chapter 4 and in

Ref. [179], the two new curves, 160CF_No2 and 250CF also have a linear relationship from

1st cycle to 10,000th cycle. As it was already established in the last chapter, the material

106
accommodates plastic strain through cyclic creep. Such cyclic creep is found to be associated

with the developments along grain boundary, i.e., grain boundary steps become more

pronounced with the increase in cyclic creep strain. The current observation further

substantiates this correlation. Moreover, it is detected that, when the cycle number was

increased to higher than 100,000 cycles and to 1 million cycles, creep rate increased and the

slope of the curves increased. This is a clear sign that, under the current testing condition,

permanent deformation in the form of cyclic creep continues with continuing cycling

indicating possible micro damage development, i.e. more severe intensification of grain

boundary extrusion.

In the last chapter, it was also seen that there were lack of PSBs formation under pure

cyclic compression conditions that were studied. To further confirm that PSB formation was

not the plastic strain accommodation mechanism, TEM and various surface examinations

were carried out to investigate. In next section the observations of dislocation morphology

are presented.

5.3. Dislocation Structure

To verify whether or not the PSBs exist, TEM was carried out to examine the

dislocation morphology in the high-stress and high-cycle conditions, i.e., 250CF after

1,000,000 cycles and 332CF after 100,000 cycles. It should be pointed out that the 332CF

condition is specifically designed to incorporate the Taylor stress plateau [9], of which the

peak stress level of 332CF equals to 98MPa. It was found in Wang and Laird‘s study [9] that

such peak stress amplitude, 98MPa, enhances the formation of PSBs and hence strain

localization in fatigued polycrystalline copper. In other words, PSBs formation, if it was

favoured, should be promoted under the application of this peak stress amplitude [9].

107
Figure 5-2: Dislocation morphology of a) 250CF after 1,000,000 cycles b) 332CF after
100,000 cycles

The TEM micrographs of 250CF (Figure 5-2 (a)) and 332CF samples (Figure 5-2 (b))

show that both exhibit apparent cell structures although not mature in their dislocation

morphology. The difference between these two conditions is that the cell structure appears to

be denser and more mature in the 332CF than that of the 250CF due to the higher stress

condition in the former. These again coincide with the observations made in our previous

work [16] that the cell structure is the resulting dislocation morphology under pure

compression fatigue, which takes form mainly in the first cycle. Thus the maturity of the cell

structure is primarily governed by the peak stress load. Nevertheless, it is seen even after

such large number of cycles and under such high load conditions, no PSBs are found in either

case.

5.4. General Surface Morphology Evolutions – Detection of Crack Nucleation

Concurrently, the same conclusion of non-PSBs existence was also drawn from the

systematic surface observations. This was first reaffirmed by the additional sets of AFM

measurements on the development of grain boundary offsets height on the 100CF sample as

shown subsequently.

108
5.4.1. AFM Observations

Using AFM results of the 1st, 100th, 5000th and 100,000th cycle from 100CF as

examples, the height profiles along a grain boundary and the corresponding grain boundary

morphology images are shown in Figure 5-3 (a)-(d) and (e)-(h) respectively. These figures

reveal the details of the surface topography evolution of the offsets due to grain boundary

extrusion. Having the ordinate being the offset height measurement under the same scale for

all measurements in Figure 5-3 (a)-(d), the profile of the offsets along the grain boundary are

shown to increase in intensity from cycle 1 to cycle 100,000. Actual numerical measurements

of grain boundary offset height and slip bands are summarized in Figure 5-4 (a) and (b),

respectively. The development of grain boundary offset height is seen to consistently

increase with the number of cycles, Figure 5-4 (a). AFM measurements on slip bands, on the

other hand, show that almost no further development beyond the first cycle, Figure 5-4 (b).

Such revelation has led to the semi-in-situ SEM observation in order to further investigate the

advancement of grain boundary extrusion at higher number of cycles.

109
Figure 5-3: AFM profile results and corresponding images at a grain boundary: (a)-(d): the profile measurements at different numbers
of cycles after 1, 100, 5000, and 100,000 cycles, respectively; (e-h) the corresponding images of the grain boundary.

110
Figure 5-4: AFM results of 100CF showing (a) grain boundary offset height measurements
vs. cycle number at 2 different locations and (b) slip band offset height
measurements vs. cycle numbers at 2 different locations

111
5.4.2. Crack Detection

Evidently the results from both TEM and AFM had shown that PSBs are not likely to

form under pure compression fatigue. The same observation was made by SEM examinations

based on the thorough studies of various peak stress conditions. In addition, parallel to the

AFM results, it was also noted from the SEM observations that only grain boundary

extrusions were prevalent under the application of cyclic compression. To fully explore the

grain boundary extrusion phenomena and the possibility of crack nucleation, the surface

morphology of the tested samples was traced periodically to higher number of cycles on

selected conditions.

Figure 5-5: Possible crack site on 120CF after 15,000 cycles: (a) low magnification 800x (b)
higher magnification 3000x

It was first observed that, for testing condition of 120CF, after the sample was cycled to

15,000 cycles, crack-like sites were noticed at grain boundaries as shown in Figure 5-5 (a) at

low magnification and (b) at higher magnification. In another testing condition, i.e. 160CF,

possible crack nucleation sites were also detected after 100,000 cycles along grain

boundaries. Figure 5-6 presents two possible crack nucleation sites under low, Figure 5-6 (a)

and (c), and high magnifications, Figure 5-6 (b) and (d), respectively. In addition, it was

seen that under 120CF condition and after 200,000 cycles, there is strong evidence of plastic

112
deformation at the grain boundaries at higher magnification as seen in Figure 5-7. Such

evidences has prompted the application of higher peak stresses for testing to the level of

250CF and 332CF, so that these surface phenomena and their development could be detected

more readily and clearly.

Figure 5-6: Two possible crack sites on 160CF after 100,000 cycles. (a) Site 1 at a low
magnification 2000x, (b) Site 1 at a higher magnification 5000x, (c) Site 2 at a
low magnification, 2000xand (d) Site 2 at a higher magnification 3000x

113
Figure 5-7: Site 1 of 120CF condition in Figure 5-5 after 200,000 cycles (a) at 8000x (b) at
20,000x at the white circular site in (a)

In Figure 5-8, a series of optical images were taken on the sample cycled at 250CF at

different number of cycles. It is evident that the sample surface also experienced the same

roughening process as in the conditions of 100CF to 160CF [179]. i.e., the sample surface

roughens as the number of cycle increases. Such roughening morphology is also determined

to be a result of mainly the grain boundary activity as presented above and the moderate slip

band formation. As it is seen in Figure 5-4, once the slip bands formed on the surface in the

1st cycle, they did not advanced in any major form the following cycles. However, the grain

boundary phenomenon, i.e., grain boundary offsets height, is seen to develop in severity as

cycle progresses.

114
Figure 5-8: Optical trace of a specific site where a surface offset is developed along grain boundaries at different number of cycles for
250CF: (a) 0 cycle (b) 1 cycle (c) 10,000 cycle (d) 150,000 cycles with recorded mean strain marked as εpl on the top left.

115
Figure 5-9: SEM trace of the site correspond to circled region in Figure 5-8 at different number of cycles: (a) 10,000 cycle @ 5000x
(b) 50,000 cycle @ 5000x (c) 150,000 cycle @ 5000x (d) 10,000 cycles @ 20,000x, (e) 50,000 cycles @ 20,000x (f) 150,
000 cycles @ 20,000x with recorded mean strain marked as εpl on the top left.

116
Figure 5-10: Crack nucleation site on 250CFand its development from (a) 50,000 cycles @ 6000x, (b) 150,000 cycles @ 6000x, and
(c) 1,000,000 cycles @ 6000x and (d) at the larger magnification at 1,000,000 cycles @ 10,000x with recorded mean strain
shown as εpl.

117
The same development in grain boundary offsets is also observed on the surface of the

250CF sample. Using Figure 5-8 (a)-(d), as a specific example, it is seen that noticeable step

already emerged on the surface after 10,000 cycles as indicated by the circled area. The

phenomenon continues to intensify as the number of cycle progresses. Figure 5-9 (a)-(f) are

the corresponding SEM micrographs that document this development. Clearly, the grain

boundary offset is seen to have gained in height from 10,000 cycles (Figure 5-9 (a) & (d)) to

150,000 cycles (Figure 5-9 (c) & (f)). Comparable to the plastic deformation noted for 120CF

along the grain boundary, the same but more severe plastic deformation morphology at the

grain boundary was seen for 250CF from Figure 5-9 (d) - (f). Another interesting

phenomenon is also observed in these images. From 10,000 cycles to 50,000 cycles, a set of

wavy grain boundary extrusions are seen to develop parallel to the grain boundary offset as

marked by the two white arrows in Figure 5-9 (e). These extrusions also become more

pronounced with increasing cycle numbers

Similar grain boundary phenomena and activities were detected in other sites as well on

the 250CF sample such as the site shown in Figure 5-10. Not only the severity of grain

boundary extrusion are seen to intensify with the cycle number from Figure 5-10 (a) to (d),

but also the extrusion in the grain interior which is parallel to grain boundary is observed to

have also further developed into an obvious crack as indicated in Figure 5-10 (d).

Furthermore, another type of extrusion morphology was also identified occasionally in

the interior of grains in both 250CF and 332CF conditions. As an example, Figure 5-11 is

one of these few sites on 332CF condition after 100,000 cycles, which demonstrates this type

of extrusion. With further EBSD analysis, the observations reveal that these extrusions

coincide with twin boundaries as shown in Figure 5-12. Figure 5-12 (a) & (b) are the SEM

image and the EBSD map of the corresponding site that is shown in Figure 5-11. With the

twin boundaries outlined in red in the EBSD map, it is observed that the extrusion is indeed

developed along the twin boundary. Figure 5-12 (c) & (d) illustrate another example of such

118
occurrence in the 250CF condition after 1,000,000 cycles. It should be noted that the

occurrence of the twin boundary extrusion was very rare. Through a large amount of

thorough examinations of various peak stress conditions, only a total of 5 twin planes were

found to exhibits such phenomenon. As such three were found in 250CF and another two in

332CF.

Figure 5-11: Twin boundary protrusion site in condition of 332CF after 100,000cycles: (a)
optical and (b) SEM micrograph of the circled region in (a) at higher
magnification @ 10,000x

119
Figure 5-12: (a) SEM image and (b) the EBSD analysis of twin boundary extrusion sites for 332CF condition after 100,000 cycles.
Note: This is the same site as shown circled in Figure 5-11. In parallel, (c) and (d) show the SEM and EBSD analysis for
the 250CF condition after 1,000,000 cycles.

120
Two types of EBSD maps were also generated to analyze the grain orientation and the

grain boundary character as well as the local plastic deformation. Figure 5-13 (a) is a Schmid

map that outlines the individual grains and their orientation with respect to the loading

direction. In the same figure, grain boundary character of each grain boundaries is also

revealed, where the twin boundaries are identified by red and the random boundaries are

marked with black. It is seen that the cracked grain boundary is characterized as a random

grain boundary. Figure 5-13 (b) is an EBSD local mis-orientation analysis that mapped out

plastic strain qualitatively. The analysis showed that the protruded grain boundaries have the

highest local mis-orientation indicating that they have sustained more plastic strain due to the

mis-orientation related deformation incompatibility.

Figure 5-13: Two EBSD Maps taken at a crack site for 250CF condition after 1,000,000
cycles (a) local misorientation map and (b) Schmid factor along with the grain
boundary character map
121
Overall it was noted from all these surface observations that, where the above extrusion

phenomena were observed, the grain boundaries, including twin boundaries, seem to have a

certain orientation relationship with the loading axes. To further analyze this relationship, the

angles between these extruded grain boundaries and the loading axis were measured. Figure

5-14 is an example micrograph of condition 120CF that illustrates how these angles were

measured. The white lines on Figure 5-14 mark the surface trace of the protruded grain

boundaries. By measuring the angle between each of the white lines and the loading direction,

an orientation angle, θ, is obtained. The corresponding measurements from Figure 5-14 are

therefore summarized in Table 5-1. The same measurements were performed on other

micrographs of different conditions. Statistical results of these measurements of 152 grain

boundary protrusion sites have been summarized as a histogram as shown in Figure 5-15.

The actual data from Figure 5-15 are tabulated in Table 5-2. It can be seen that the overall

trend for these extruded grain boundaries, including twin boundaries were mostly at an angle

between 60° and 90° against the loading axes. Furthermore, Figure 5-15 also demonstrates

that as the angle between the grain boundary and loading axis increases, the possibility for

grain boundary extrusion also increases. i.e., the potential for a grain boundary forming grain

boundary step is higher when the angle between the grain boundary and the loading axis is

high. Such trend peaks when the angles are within the range of 80° to 90°. It should be noted

that amongst the 39 angles that are between the 80° to 90°, there are 12 grain boundaries,

which are at 90° with the loading axis.

122
Figure 5-14: Illustration of grain boundary angle measurements at grain boundary protrusion
sites for condition 120CF. A total of 9 protruded grain boundaries angles, 
against the loading axis were measured.

Table 5-1: List of Grain Boundary Angle Measurements from Figure 5-14

Site 1 2 3 4 5 6 7 8 9
Orientation 60° 80° 80° 90° 60° 90° 55° 60° 80°
Angle (θ)

Figure 5-15: Distribution of protruded grain boundary angle with the loading axis.

123
Table 5-2 Summary of θ Measurements at Grain Boundary Extrusion Sites from Figure 5-15

Orientation 20° 25° 30° 35° 40° 45° 50° 55° 60° 66° 70° 75° 80° 85° 90°
Angle
Frequency 3 3 2 3 9 13 13 9 22 6 23 7 20 7 12
Total Grain Boundary Extrusion Sites = 152

5.5. Discussion

It has been long recognized that the PSBs and/or the impingement of PSB on GB are the

preferential sites for crack nucleation under symmetrical fatigue in a surface defect free

material [19, 106, 117, 171-178]. However, under the presence of a mean stress, it was also

revealed that the formation of PSBs may be suppressed when cell structures prevails such as

shown by Lukas et. al.,[75], when the condition was biased with the positive mean stress. In

the last chapter, it was already established that the cyclic deformation response of

polycrystalline copper under pure compression fatigue was mainly in the form of cyclic creep

as it was shown in Figure 5-1 along with two other new conditions. This creep behaviour was

identified to be associated with the grain boundary activities rather than the conventional

PSB development [13, 179]. In this chapter, it is again confirmed that there is indeed a lack

of strong slip behaviour and hence the absence in PSB formation, Figure 5-4. In addition, as

shown in Figure 5-2, it is seen that loosely formed cell structure is the prevalent dislocation

morphology under the sole influence of compressive stress spectrum. Different from that

reported in Ref. [75], the cell structure observed in the present result, is not as well

established or mature, even when applied stress was very high. It is also determined that

there is no trace of PSBs on the surface even at the two extreme peak stress conditions,

250CF and 332CF. This result is proven to be more significant as one of the high peak stress

load conditions is equal to the Taylor stress plateau stress. These observations indicate

convincingly that, if fatigue cracks would initiate under the present testing condition, it

would not be mainly due to PSB- related protrusion and intrusion.

Furthermore, unlike the study reported in Ref. [75] with single crystals, the current
124
investigation was carried out with polycrystalline copper samples, the sliding of a crystal

slab on a specific slip plane was not possible due to its interactions with adjacent grains and

was indeed not detected through rigorous surface observation. Instead, the observed

phenomena show that the fatigue crack nucleation process under pure compression fatigue is

merely the development of grain boundary step that leads to the crack nucleation. The grain

boundaries that formed steps are identified to be high angle grain boundaries and they are

oriented with surface trace between 60° - 90° to the loading axis. Furthermore, the most

vulnerable grain boundaries are identified to be 80° - 90° against loading axis. Such

observation has also been confirmed similar to that reported by Kim and Laird [116] but

under different fatigue loading conditions. In Ref. [116] the grain boundary step formation

under high strain fatigue condition was due to the asymmetrical slip activities between

adjacent grains. The continuing grain boundary step growth was mainly contributed by cross

slip in their tested condition. However, the same grain boundary extrusion mechanism could

not be said for the present investigation due to the lack of intensive slip behaviour.

As it was shown in the previous results with notched specimens [13], fatigue cracks may

nucleate when the stress concentration reaches certain magnitude. It was further revealed that

cracks nucleate where maximum shear stress was favoured. In the present study, it is seen

that grain boundary extrusion is the sole active mechanism for accommodating permanent

strain that creates surface roughness. Thus, although there are no surface irregularities in the

subject specimen, stress concentration sites could still arise from grain boundaries. Margolin

and his colleagues [180, 181] have demonstrated that localized stress fields at the grain

boundaries was always at the highest compared to the grain interiors. Due to the different

orientation between two adjacent grains, there would be different amount of deformation in

each grain, hence giving rise to the elastic distortion at the grain boundary regions.

Subsequently if the grain boundary was oriented as such that it contains the maximum shear

stress direction towards the surface, it would prompt for the flow of material towards the

125
surface leading to the formation of grain boundary step.

Figure 5-16 is a schematic figure demonstrating the orientation relationship between a

grain boundary and the loading axis on the sample surface. As an example, Figure 5-16 (a)

shows a grain boundary plane with 3 different orientations (P1, P2 and P3) and all three

arrangements possess the same surface trace at a 90° angle with the loading axis. Although

the plane may take on different tilt angles in the depth direction, the maximum shear is

always found on the plane that has the 45° tilt towards the surface (P2). Therefore, despite

the grain boundary appearing to be perpendicular to the loading axis from the surface view,

the tendency for surface step forming may still be high as revealed by the histogram in Figure

5-15. However, when the grain boundary angle on the surface with the loading axis starts to

deviate from the right angle while the plane normal is still 45° against the loading axis, plane

P1 in Figure 5-16 (b) shows that the maximum shear direction is no long towards the free

surface. This is revealed when comparing P1 and P2 in Figure 5-16 (b), where grain

boundary plane (P2) in Figure 5-16 (b) is the same grain boundary plane (P2) shown in

Figure 5-16 (a). It demonstrates that the maximum shear stress flow directions of the two

planes are different as indicated by the black dash arrows. Since the grain boundary plane (P1)

is no longer perpendicular to the loading axis, the maximum shear flow directions shifted

away from being towards the surface to the neighbouring grains. In this instance, the surface

trace angle of grain boundary plane (P1) is at 70° to the loading axis. Hence, the possibility

of surface step forming at such angle is reduced as seen in the histogram in Figure 5-15. This

is even more so as the grain boundary surface trace angle becomes even smaller against the

loading axis and thus lowering tendency of step formation at these grain boundaries.

Nevertheless, once the grain boundary steps are extruded by the maximum shear stress, it is

expected there would be high strain field concentrations at these active grain. As it is shown

from the EBSD local misorientation map, Figure 5-13 (a), high localized plastic strains is

indeed observed at these grain boundaries.

126
Figure 5-16: Illustration of different cases of grain boundary orientation in relation to the
specimen surface and to the loading axis when (a) A grain boundary has its
surface trace perpendicular to the loading axis but it may tilt at different angles in
the depth direction, and (b) Grain boundaries that have different surface trace
angle to the loading axis. P1 is at 70 o, where P2 is at 90o but both at the same tilt
angle of 45° in the depth direction.

Moreover, depending on the relative size of the grains that are joined on either side of a

grain boundary, not only the stress concentration at this grain boundary may be changed, but

also there will be varying degree of deformation incompatibility effect between the grains.

When a grain boundary is joined by a large grain and a small grain, the stress concentration is

higher in the large grain than the scenario whereby a grain boundary is joined by two

relatively equal size grains. As the smaller grain usually encounters more deformation

constraint due to its bigger surface to volume ratio, i.e. the smaller grain would experience
127
stronger incompatibility effect, the tendency for the small grain to be deformed would be

weaker and thus the propensity for large grain to be deformed is higher. This was clearly

observed in the present study. As demonstrated by Figure 5-9 and also Figure 5-12 (b), where

the grain boundary is joined by a large and a small grain and the cracked grain boundary is

on the large grain side.

For the extrusion-like-features at twin boundaries, analogous to the existence of elastic

distortion at the grain boundary region, they are also explained by a similar mechanism. It

was determined by Wang and Margolin [122] from their FEM analysis that when the twin

bicrystal is subject to the applied load, regardless whether the applied stress is tensile or

compressive, the local increase in elastic stress along the twin plane is the highest or

considerably enhanced. It is due to the inhomogenous deformation between the two sides of

the twin boundary that give rise to the slipping of the boundary plane in FCC crystals and

hence yielding the extrusion-like feature on the surface, as shown in Figure 5-12.

Nevertheless, the extrusion-like feature formed at twin boundaries should not be mistaken as

the PSBs in the present case.

By nature pure copper has relatively high stacking fault energy, which promotes cross

slip, hence it is in favour of PSBs formations. However, the current condition shows that due

to the strong deformation incompatibility between the adjacent grains, it has limited the

number of operating slip systems to one. In other words, the operations of other slip systems

are subdued. As PSBs requires multiple slip systems to form, the current conditions do not

favour for such formation. Thus, even for a wavy-slip material like copper, the cracks are

nucleated like a planar slip material instead of the conventional PSBs formations.

Overall the results in this chapter have further substantiated the correlation of cyclic

deformation under pure compressive load, i.e. cyclic creep, and the development of grain

boundary steps as it is seen from Figure 5-1 and Figure 5-9. The cyclic creep rate at the

initial 104 cycles reflects the formation of the apparent grain boundary step as demonstrated

128
in Figure 5-9 (a) at low magnification, or Figure 5-9 (d) at high magnification. As cycling

continues from 104 cycles and onwards there will be changes in cyclic creep rate and

experimental results show that there is an intensification of grain boundary step as observed

in Figure 5-9 (b and c) to (c) at low magnification or Figure 5-9 (e and f) at high

magnification. Such intensification and changing in cyclic creep rate may indicate crack

nucleation process. In the 250CF condition, grain boundary extrusion intensity is reflected by

the introduction of the parallel steps which leads to the eventual crack nucleation. In the case

of 120CF and 160CF, although the obvious parallel steps were not seen on the lower

magnification, the same intensification process show that it did occur along the grain

boundary which then leads to crack-like sites on the 120CF and 160CF sample.

Finally, if the current existing microcracks were to continue experiencing the pure

compressive load, the possibility for further development may be viewed as less likely.

Given the material's high ductility, the propagation of this form of crack may be slow and

less likely to develop into a major crack. However, if the applied load were to reverse to

tensile, depending on the magnitude of the applied load, these microcracks may continue to

grow.

5.6. Summary

As a further investigation of cyclic deformation phenomena under pure compression

fatigue, cyclic deformation behaviour at higher loading condition and longer cycles has been

studied. The possibility of crack nucleation has also been verified and thus the crack

nucleation mechanisms have been determined. The major conclusions are:

(i) Crack nucleation is proven to be possible under pure compression fatigue even

when there is no surface irregularity. Such crack nucleation mechanism is still

associated with certain form of localized stress concentration site, i.e., grain

boundaries. Hence crack nucleates through the grain boundary extrusions.

(ii) The grain boundary extrusion phenomenon results from the strong deformation

129
incompatibility at the grain boundary, whereby grain orientation and relative

size play important roles in influencing which grain boundaries to form grain

boundary steps.

(iii) The grain boundary extrusions are formed by maximum shear at these high

angle boundaries or twin boundaries, which have a surface trace of 60°- 90°to

the loading axis. In addition, cracks are formed at the strongest grain boundary

incompatibility regions where there is high localized stress concentration, i.e.,

between large and small grains.

(iv) Cyclic compression does not yield PSBs even under conditions that were

favourable for the formation of PSBs under symmetrical fatigue but result in

grain boundary extrusion. Consequently cyclic creep is strongly dependent on

grain boundary activities.

(v) The change in cyclic creep rate is attributed to the micro crack nucleation

process after 104 cycles.

(vi) Cell structure, in spite not well developed, is indeed the prevailing dislocation

morphology under pure compression fatigue regardless of the peak stress

conditions.

(vii) Statically, the majority of grain boundary extrusions occur at random grain

boundaries as opposed to twin boundaries.

130
6. Results and Discussion III:
Asymmetrical Mechanical Response of OFHC Polycrystalline Cu

6.1. Overview

Cyclic deformation responses of OFHC copper under pure compression fatigue

conditions have been discussed in detail in the previous chapters. Comparison between these

responses under pure cyclic compression conditions and those under symmetrical fatigue

conditions has also been presented in these chapters. It was revealed that the overall cyclic

deformation response of OFHC copper under pure compression cycling condition exhibits

rather dissimilar behaviour compared to those under general push-pull fatigue conditions.

The same conclusion was also drawn from the contrast made between the present studies and

that of monocrystalline copper under pure tension fatigue. As it was seen from the results of

uniaxial tensile and compression tests, there was discrepancy in yield strength between

tensile and compression tests. This may imply that there exists certain degree of asymmetry

between opposite loadings. According to Laird [182], asymmetric behaviour between tension

and compression was indeed observed in plastic strain controlled symmetrical fatigue

conditions for monocrystalline Cu, the emphasis of this chapter is therefore placed on the

examination of polycrystalline OFHC Cu in terms of its symmetry between tension and

compression in load controlled symmetrical fatigue conditions. In addition, since

Bauschinger effect describes the asymmetrical mechanical behaviour, Bauschinger

parameters are also evaluated.

6.2. General Cyclic Deformation Response of Cu under Symmetrical Conditions

When symmetrical fatigue was carried out to investigate the difference between

compression fatigue and symmetrical fatigue conditions, an interesting phenomenon was

observed from the summary of the symmetrical hysteresis loops. Figure 6-1 and Figure 6-2

show the summary of hysteresis loops obtained from two different peak stress conditions,

100% σy and 120% σy for symmetrical fatigue conditions, respectively. In each of this peak

131
stress conditions, the loading spectrum may be tension-started followed by compression load

or vice versa compression-started then reversed in tension so the results are classified as TCF

or CTF respectively.

From Figure 6-1 and Figure 6-2, the general trend of these hysteresis loops summaries

clearly indicates that there is a large net unidirectional strain accommodated in the first half

cycle despite the loading condition is symmetrical. In other words, the mean strain does not

equal to zero even when the applied load spectrum was symmetric in tension and

compression. Figure 6-3 shows further analyses of mean strain with respect to the number of

cycles for each peak stress conditions. It is demonstrated that a unidirectional strain is always

introduced in the 1st cycle for all of the peak stress conditions. However, such unidirectional

strain dropped in magnitude after 2nd cycle and maintained a steady mean strain for the

remaining cycle span. It was also found that depending on whether the first half cycle is

tensile or compressive, i.e., tension-start (TCF) or compression-start (CTF), the

unidirectional strain always retains the same polarity as introduced in the starting stroke.

It should be noted that such an observation was rarely reported as most of the studies on

cyclic deformation only describes the plastic strain evolution not the summary of the entire

hysteresis loop. Nevertheless, Weiss et al. [22] have documented the same mean strain shift

as shown in Figure 6-4; where they investigated the influence of different ramp loading

period and the effect of various ramp amplitude of polycrystalline Cu. It was reported that

regardless of the ramp loading conditions, there is always build-up unidirectional strain. The

build-up unidirectional strain was especially noticeable as in Figure 6-4 in condition of which

ramp loading was not applied, i.e., the condition that is comparable to the current TCF

condition. In the same reference [22], it was also mentioned that the unidirectional strain was

associated with the first cycle strain. Such commentary has signified the importance of the

first cycle. Hence, the emphasis is placed on the first cycle hereafter. Bauschinger effect from

first cycle of different peak stress and first stroke conditions are also evaluated.

132
Figure 6-1: Summary of hysteresis loops for peak stress condition of 100% σy for (a) tension-
start (100TCF) and (b) compression-start (100CTF)

133
Figure 6-2: Summary of hysteresis loops for peak stress condition of 120%σy for (a) tension-
start (120TCF) and (b) compression-start (120CTF)

134
Figure 6-3: Mean strain analysis of selected conditions plotted against cycle number

Figure 6-4: Mean strain analysis of selected conditions plotted against cycle number from

Weiss et al. [22]

135
6.3. Evaluation of Bauschinger Parameters

In this section, Bauschinger effects are evaluated for the first cycle using five different

parameters from Figure 2-27, which are evaluated based on three different peak stress

conditions presented in Figure 6-5, Figure 6-6, and Figure 6-7. Since for copper, the material

flow occurred well before the conventional engineering yield point (σy at ε = 0.2%), a new

yield criterion is needed. The yield points determined for Bauschinger effect at the forward

and reverse load were both defined as ε = 0.025%. Therefore, the parameters used to analyze

the Bauschinger effect are forward yield stress (σy,0.025%), reverse yield stress (σR, 0.025%),

permanent softening (∆σP), forward strain (εP), and reverse strain(εR), where εP and εR are

εPl,tension and εPl,compression, respectively. Table 6-1 summarizes these parameters from Figure

6-5, Figure 6-6, and Figure 6-7. It should also be emphasized here that the peak stress

conditions for all the tests are weighted with the conventional yield stress (σy) at ε = 0.2%

and not based on the new definition.

Table 6-1: Summary of Bauschinger Parameters for Different Peak Stress Conditions

TCF Conditions CTF Conditions


90TCF 100TCF 120TCF 90CTF 100CTF 120CTF
σy,0.025%: 19.4 20.5 20 24.1 25.5 23.9
σR, 0.025%: 23 23.8 23.4 17.4 19.9 21.9
∆σp: 1.8 1.7 3.3 3.8 4.32 4.8
εp: 0.0011 0.0016 0.0043 0.0014 0.002 0.0052
εR: 0.0006 0.0008 0.0016 0.0016 0.0018 0.0022
β=εR: 0.0006 0.0008 0.0016 0.0016 0.0018 0.0022
β/ εp: 0.55 0.50 0.37 1.14 0.90 0.42

136
Figure 6-5: Bauschinger effect evaluation for peak stress condition of 90% σy for (a) tension-
start (90TCF) and (b) compression-start (90CTF)

137
Figure 6-6: Bauschinger effect evaluation for peak stress condition of 100% σy for (a)
tension-start (100TCF) and (b) compression-start (100CTF)

138
Figure 6-7: Bauschinger effect evaluation for peak stress condition of 120% σy for (a)
tension-start (120TCF) and (b) compression-start (120CTF)

139
To identify the difference between the two different starting stroke conditions as well as

to reveal the Bauschinger effect for each condition, several comparisons are made based on

these parameters. When all the yield strengths obtained from the forward (σy,0.025%)and

reverse (σR,0.025%) of all peak stress conditions are plotted on the same graph as shown in

Figure 6-8, these yield strengths, as expected, are seen to be independent of the peak stress

conditions. On the other hand, when the forwarding yield strengths obtained for TCF

conditions are compared with that of CTF, it is evident that the σy,0.025% of CTF is always

higher than the σy,0.025% of TCF. However, the trend is reversed upon load reversal. The

σR,0.025% of TCF is always higher than the σR,0.025% of CTF.

Figure 6-8: Yield strengths comparisons

Dissimilarities between the TCF and CTF condition were also evident through the

analysis of Bauschinger effect. As reviewed in Chapter 2, Bauschinger effect of the first

cycle is revealed simply by comparing these parameters from the forward half cycle and the

140
subsequent reversed half cycle. For TCF conditions, they have collectively showed that there

seem to be an ―anti-Buaschinger‖ effect or ―reversed Bauschinger effect‖. It was observed

that instead of lowered yield stress, an elevated yield was obtained, i.e., σy,0.025% < σR,0.025%. In

contrast, for CTF conditions, they have demonstrated that conventional Bauschinger effect is

always experienced, i.e., σy,0.025% > σR,0.025%. However, when the permanent softening (∆σP)

is evaluated, a contradicting result has been identified. Instead of an anti-Bauschinger effect

as characterized by the forward and reverse yield strengths comparison, TCF shows that there

is still a small work hardening effect in all three conditions. i.e., ∆σP > 0. These magnitudes

of ∆σP in TCF conditions were also comparatively small to that of the equivalent CTF peak

stress conditions, which implies the first tension-stroke in the TCF condition has introduced

much less back stress compare to that of first compression-stroke in CTF, a major

microstructural reason responsible for Bauschinger effect.

Figure 6-9: Bauschinger effect evaluation for peak stress condition of 120%σy for (a) tension-
start (120TCF) and (b) compression-start (120CTF)

141
On the other hand, upon analyzing the Bauschinger effect using the Bauschinger strain

parameter β, it is seen from Figure 6-9 that the ratio of (β/ εp) in the TCF conditions is always

smaller than the CTF conditions. The relative small ratio of (β/ εp) in the TCF conditions

signifies that the materials under TCF conditions always work hardened more. TCF

conditions, in other words, displays much larger Bauschinger effect than the CTF conditions.

Notably from all the Bauschinger effect assessments, not only the difference in Bauschinger

behaviours between TCF and CTF conditions have been realized, but also the contradiction

within each individual condition has been seen in terms of the Bauschinger effect

establishment. Specifically, the evaluation of Bauschinger effect shows that using different

Bauschinger parameters and different comparison may lead to different Bauschinger effect

interpretations. As a result, further analysis is required to explain these inconsistencies.

Figure 6-10: Bauschinger effect evaluation for peak stress condition of 120%σy for (a)
tension-start (120TCF) and (b) compression-start (120CTF)

Moreover, another interesting discovery was also made from the CTF conditions where

142
the shape of the compression-half cycle versus the tension-half cycle was found to be

dissimilar. Comparing the two initial half strokes of the first cycle from peak stress condition

of 120%σy as seen in the red circle in Figure 6-10, the compression curve has a sharper

elastic to plastic transition in the curve than a more gradual one of the TCF curve. This

phenomenon does not restrict to only one peak stress condition but apply to all three and will

be referred as the ―sharp yielding phenomenon‖ hereafter. Due to such ―sharp yielding

phenomenon‖, the forwarding yield strength is always higher in the CTF conditions than that

of the TCF conditions. Furthermore, since similar observation is not noted in the second

compression half stroke in all TCF condition in Figure 6-5, Figure 6-6, and Figure 6-7 , such

compression sharp yielding phenomenon is concluded to occur only in the fresh sample

during first testing. Thus, tensile and compression curves were re-examined from the static

uniaxial tensile and compression tests.

6.4. Mechanical Properties of Cu

As the engineering stress strain curves of polycrystalline OFHC Copper are re-examined,

it is also observed that the stress-strain profile at the transition point from elastic to plastic

region between the tensile and compression curves was apparently different. Similar to the

compression curve from the compression-start symmetrical fatigue conditions, such

difference is revealed in the red circle in Figure 6-11, when the two curves are plotted up to

εeng = 0.3%. Despite the two curves share the same elastic modulus up to σ 10MPa, i.e., the

same slope in the elastic region, and the similar hardening rate in the plastic region, the

compression curve has a drastic turn from elastic to plastic compare to the moderate

transition observed for the tensile curve.

143
Figure 6-11: Comparison of engineering stress strain curves between tensile and
compression tests

6.5. Explanations for Discrepancies in Bauschinger Effect between Two Loading

Conditions

In an effort to provide the explanations to the inconsistences that were observed within

the two conditions from the investigation of Bauschinger effect, further analysis has been

carried out. Strain hardening rate analysis was first conducted since it was apparent that the

strain hardening behaviour is quite different between tension and compression. As it was

noted from both the uniaxial mechanical experiments and the symmetrical fatigue tests that

the hardening behaviour of both TCF and CTF cases before σ 10MPa appears to be very

similar, the strain hardening rate from this stress point on is therefore specially studied for the

symmetrical fatigue conditions. In addition, since strain hardening is generally associated

with dislocation activities, dislocation morphologies were also examined in this section.

6.5.1. Analysis of Strain Hardening Rate

Figure 6-12 plots the strain hardening rate as function of the true strain for conditions of

144
100TCF and 100CTF. These plots are representative of the strain hardening behaviour

between the TCF and CTF conditions. The same plots for other conditions are provided in

Appendix A. The analysis demonstrates that the strain hardening rate is always elevated

when the stroke is reversed for TCF condition. It was also seen that in the same condition,

the strain hardening between the two half curves are very comparable in terms of the strain

hardening behaviour. Both show that the change in strain hardening behaviour was very

gradual. Unlike the CTF condition, the difference in strain hardening rate between the two

half curves are very distinctive. The strain hardening behaviour for compression-stroke has a

very abrupt change at approximately εtrue = 0.02%, which corresponds to the ―sharp yielding

transition‖ that was observed on stress-strain curve. The strain rate for reversed stroke on the

other hand continuously decreases until the two strain hardening curves converge.

In addition, a very dissimilar strain hardening behaviour between the two forward

strokes is clearly seen. The strain hardening is always higher in the TCF condition than that

of compression start conditions after the sharp yielding phenomenon. However, the strain

hardening rates for the two conditions shows that convergence may be reached at the higher

strain. Such convergence is predicted to be close to (ε = 0.16%). In other words, the two

forward curves share the same strain hardening rate after the conventional yield. This was

indeed observed in the uniaxial tensile and compression curves as presented in Figure 6-11.

145
Figure 6-12: Strain hardening rate as a function of true strain for conditions (a) 100TCF (b)
100CTF

146
Figure 6-13: Comparison of strain hardening rate for the forward stroke between 100TCF
and 100CTF

Evidently, the observed inconsistencies from the Bauschinger effect assessments are due

to the dissimilarity in strain hardening behaviour in the two conditions. For instance the

observation of ―anti-Bauschinger effect‖ from the TCF may be explained by the variance in

strain hardening rate at the initial portion for each half strokes in TCF. As it can be seen from

Figure 6-12 (a), the strain hardening rates of the reversed compression stroke is always much

higher than the forward stroke prior to εtrue = 0.025%. Therefore, a higher yield from the

reversed stroke can be expected. In contrast, for CTF condition in Figure 6-12 (b), at the εtrue

near 0.025% it can be seen that the strain hardening rates of compression forward stroke is

higher than the reverse load and hence a conventional Bauschinger effect is observed. Such

effect may be easily reflected upon by the shape difference in the compression stress strain

curve vs. that of tensile as shown in either Figure 6-10 or Figure 6-11. Since the existence of

―sharp yielding phenomenon‖ in the compression curve versus the more rounding shape of

147
the tensile curve, lowering yield stress is likely to occur.

The explanation offered for the observed diverging trend of ∆σP between TCF and CTF

is also no different than the above discussion. As it could be demonstrated from Figure 6-12

(a) that at the strain point where ∆σP was evaluated, the corresponding strain rate for each

forward and reverse loading curve are drastically different for TCF. The strain hardening rate

of the reverse curve was much larger than the strain rate forwarding curve at the strain point

where ∆σP was evaluated, the ∆σP magnitude should be very small. By the same analogy, for

CTF condition, since the strain hardening rates at which ∆σP was taken appears to be

comparable for both forwarding and reverse curve, ∆σP is shown to be relatively larger.

On the other hand, as the hardening rate of tension stroke in TCF is always higher than

the compression stroke in CTF, the material under TCF condition would experience higher

hardening effect than when it is subjected to the CTF condition. Therefore, there will be less

hardening strain recovery during the reversal loading for the TCF condition than in the CTF

condition. This ultimately explains why when the Bauschinger effect was evaluated using (β/

εp), the CTF always displays a much larger ratio than that of TCF conditions.

Based on the above analysis and observations, it is clear that the behaviour of tension

and compression is highly asymmetrical. The difference in Bauschinger effect is therefore

the results of such asymmetrical behaviour.

6.6. TEM Observations

To further investigate the cause for the difference in strain rate, dislocation behaviours

was examined for two peak stress conditions using TEM. Figure 6-14 shows four TEM

images, which includes the conditions of (a) 90TCF, (b) 90CTF, (c) 120TCF and (d) 120CTF.

It is seen from both Figure 6-14 (a) and (b) there are some lose dislocation networks for peak

stress condition of 90% σy. However, upon comparing the two TEM micrographs, it is

observed that the dislocation structures in the two conditions are very similar. The dislocation

density is almost equivalent. Similar observation was also made upon the comparison

148
between the 120TCF and 120CTF. The TEM micrographs in Figure 6-14 (c) and (d) has

showed that despite the dislocation density increases from the 90% σy peak stress conditions,

the dislocation structures remains very similar between the two different applied initial stroke

conditions in terms of its density and pattern.

The aforementioned discovery has suggested that the difference in hardening rate

between the two conditions, TCF vs. CTF, was not the result of dislocation activities. Such a

result has raised a significant issue: what other fundamental mechanisms could give rise to

the asymmetrical behaviour between tension and compression for polycrystalline OFHC Cu?

According to Weiss et. al.[22], tensile slip bands were produced most profusely in their

tested condition when ramp loading was zero, of which the condition is equal to our TCF

condition. If so, would the same slip bands be produced if the condition starts with

compression load? Perhaps the difference in fundamental mechanisms between tension and

compression, which lead to this asymmetrical behaviour, is the main reason for the overall

difference between compression fatigue and the tension fatigue. In order to answer these

questions, future work is subsequently needed.

149
Figure 6-14: Dislocation structure of conditions (a) 90TCF (b) 90CTF (c)120TCF and (d)
120CTF

150
6.7. Summary

In this chapter, behaviour of polycrystalline OHFC Cu samples under symmetrical

fatigue condition has been studied with two different starting stroke conditions, i.e., TCF vs.

CTF under three peak stress conditions. The emphasis has been placed on examining the

symmetry between tension and compression in load controlled symmetrical fatigue

conditions. In specific the first cycle of each is analyzed and characterized using Bauschinger

parameters. The major findings and observations from this investigation are summarized as

follows.

(i) Overall, polycrystalline OFHC Cu exhibits asymmetrical behaviour between

tension and compression.

(ii) Symmetrical fatigue conditions produce unidirectional strain regardless of the

starting condition, mainly in the first cycle.

(iii) The polarity of the unidirectional strain is dependent on the first stroke.

(iv) Behaviour of first cycle from tension-start (TCF) and that of compression-start

(CTF) have shown to be very different from the perspective of Bauschinger

effect. Such different behaviour was found to arise essentially from the

dissimilar strain hardening in the initial plastic deformation behaviour between

tension and compression.

(v) The difference in strain hardening behaviour between the two conditions, TCF

vs. CTF, does not seem to be the result of dislocation activities.

151
7. Conclusions
There are cases that show cracks may nucleate under pure compression fatigue

conditions such may be as a major application as the landing gear shock strut. However, as

the subject of pure compression fatigue has always been overlooked for the assumption that

pure compressive fatigue or the compression portion of the load spectrum does not promote

fatigue crack initiation or propagation, there is a need to establish the underlying

micromechanism under pure compressive cyclic conditions. This study systematically

investigated cyclic deformation behaviour of polycrystalline OFHC Cu under pure

compression fatigue conditions. It is the first study to detail and establish the cyclic

mechanical deformation response with the underlying microstructure evolution under such a

fatigue condition.

7.1. Summary of the Key Experimental Observations

In order to fully understand the true underlying micro-mechanisms on how crack

nucleation under fatigue from the material‘s perspective, crack deformation response and

crack initiation mechanisms of a smooth polycrystalline OFHC Cu under various pure

compression fatigue conditions have been investigated. The possibility of crack initiation

under pure compressive fatigue condition has also been ascertained. In addition, preliminary

investigation has been carried out to examine the asymmetrical response of polycrystalline

copper between tension and compression. Although summaries have been presented at the

end of each result and discussion sections, the key experimental observations of this study

are presented here.

(1) Under pure compression fatigue condition, cyclic creep was observed to be the

major form of plastic strain accommodation. There was no major cyclic plastic

strain accommodation detected, even if when the applied stress levels were

high, during such a fatigue condition.

152
(2) The cyclic creep was found to associate with the grain boundary extrusion on

the surface instead of the typical dislocation-slip activities. As such grain

boundary extrusions advances as the number of cycle increases.

(3) Although cyclic saturation was observed, the saturated cyclic plastic strain

amplitude is found to be much lower than that of the symmetrical fatigue.

There is no plateau of the CSSC under pure compression fatigue. Most

interestingly, no PSBs were detected under the current testing conditions.

(4) The testing material was found to be hardened within 10 cycles when cycled

under pure compression loads, unlike its behaviour in push-pull fatigue with the

same stress amplitude.

(5) Cyclic compression deformation results in dislocation cell structures for all

peak stress conditions tested, which take their form mainly in the first cycle.

Thus, under the testing condition, the maturity of the cell structure is primarily

governed by the peak stress load.

(6) Crack nucleation is verified to be possible under pure compression fatigue

through grain boundary extrusions.

(7) The grain boundary extrusions are formed by maximum shear at these high

angle boundaries or twin boundaries, the majority of which have a surface trace

of 60°- 90°to the loading axis. In addition, cracks are formed at the strongest

grain boundary incompatibility regions where there is high localized stress

concentration, i.e., between large and small grains.

(8) Symmetrical fatigue conditions produce unidirectional strain irrespective of the

starting condition. The polarity of the unidirectional strain is dependent on the

first stroke.

(9) The behaviour of the first cycle from tension-start (TCF) and that of

compression-start (CTF) have shown to be very different from the perspective

153
of Bauschinger effect. Such different behaviour was found to arise essentially

from the dissimilar strain hardening behaviour between tension and

compression. Similar strain hardening behaviour difference is also found in

uniaxial tension and compression.

(10) The difference in strain hardening behaviour between the two conditions, TCF

vs. CTF, was not a result of dislocation activities.

7.2. Major Scientific Contributions

Analysis of this study has yield a new understanding of cyclic deformation under pure

compressive loading spectrum from the fundamental perspective. The following conclusions

can be drawn.

(1) The overall cyclic mechanical deformation response and microstructural

evolutions of OFHC copper under pure compression cycling condition exhibits

rather dissimilar behaviour compared to those under general push-pull fatigue

or those under pulsating tension fatigue.

(2) The dissimilarity in the cyclic mechanical deformation response between

compression fatigue and conventional push-pull fatigue essentially arise from

the difference in microstructural evolution. (a) Compression fatigue does not

promote dislocation activities as opposed to the active dislocation evolution

that is observed in the conventional push-pull fatigue. (b) Compression fatigue

results in dislocation structure cell and grain boundary extrusions as opposed to

the ladder dislocation structure and the PSB formation under the stress

condition that PSB formation is favoured. Therefore a difference in CSSC is

also seen.

(3) Crack nucleation is verified to be possible under pure compression fatigue

even when there is no surface irregularity. Such crack nucleation mechanism is

154
still associated with certain form of localized stress or strain concentration sites,

i.e., grain boundaries. Hence crack nucleates through the grain boundary

extrusions.

(4) Polycrystalline OFHC Cu exhibit asymmetrical behaviour between tension and

compression. As such the Bauschinger effect of first cycle from tension-start

(TCF) and that of compression-start (CTF) have shown to be very different due

to the difference in strain hardening rate. The strain hardening difference was

also found between the uniaxial tension and compression results. However,

dislocation activities were not the micro-mechanisms for such difference in

strain hardening behaviour.

155
8. Recommendation for Future Work
The cyclic behaviour of polycrystalline OFHC Cu under pure compression fatigue has

been studied and discussed extensively in this dissertation. Key differences between

compression fatigue and that of conventional push-pull fatigue has been uncovered. New

knowledge towards completing the understanding of cyclic loading in wide range application

has also been established. However, further investigation is required to reveal deeper

information and underlying mechanisms regarding phenomena that were observed. To do so,

some of the future works are recommended here:

(1) The grain boundary extrusions were found to form by maximum shear at these

high angle boundaries or twin boundaries. To study the underlying

micromechanisms that lead to such grain boundary extrusions, further

experimental work is required to re-examine the vertical cross-sectional

morphology of these compressively fatigued samples. Through the use of

various microscopies, the observations should reveal grain boundary extrusion

morphology and the behaviour of the neighbouring grains. Finite element

analysis should also be performed according to the observation to characterize

the neighbouring interactions.

(2) The strain hardening behaviour between tension and compression was very

distinctively different. To determine the underlying cause for this dissimilar

behaviour, systematic surface observation as well as TEM observation should

be carried out. Specifically what is the mechanism behind the ―sharp yielding

phenomenon‖ for the compression curve?

(3) In order to investigate the difference in the deformation behaviour, finite

element analysis on the behaviour of uniaxial tension and compression as well

as cyclic tension and compression should be studied. Models with various bi-

crystal arrangements should be utilized.

156
(4) Investigation of symmetrical fatigue conditions for both tension-start (TCF)

and compression-start (CTF) at higher peak stress application is of an interest.

Whether or not this may emphasize more of the effect of first cycle on the

subsequent cycles when the peak stress condition is high needs further

examination.

157
9. References
[1] G.S. Campbell, A note on fatal aircraft accidents involving metal fatigue, 3 (1981) 181–

185.

[2] G.E. Dieter, Mechanical Metallurgy, in, McGraw-Hill, United States, 1986.

[3] J. Schijve, Fatigue of aircraft materials and structures, 16 (1994) 21–32.

[4] A. Levin, Holes in planes rare but pose deadly risk, in, USA Today, USA, 2011.

[5] Numerous examples of fatigue failurse are given in Failure Analysis and Prevention, in:

Metal Handbook, American Society for Metals, Ohio, 1975.

[6] D.K. Holm, A.F. Blom, S. Suresh, Growth of cracks under far-field cyclic compressive

loads: Numerical and experimental results, Engineering Fracture Mechanics, 23 (1986) 1097-

1106.

[7] ASTM Standard E647-13a, "Standard Test Method for Measurement of Fatigue Crack

Growth Rates", in, ASTM International, 2014.

[8] N.S. Currey, Aircraft Landing Gear Design - Principles and Practices, American Institute

of Aeronautics and Astronautics, 1998.

[9] M. Roth, M. Yanishevsky, P. Beaudet, Failure Analysis of Aircraft Landing Gear

Components, in: First International Conference on Failure Analysis, ASM International,

1992.

[10] R.P. Hubbard, Crack Growth Under Cyclic Compression, Journal of Fluids Engineering,

91 (1969) 625-630.

[11] W.-Y. Chu, C.-M. Hsiao, T.-H. Liu, Fatigue under cyclic compressive load, Fatigue &

Fracture of Engineering Materials & Structures, 7 (1984) 279-284.

[12] T. Christman, S. Suresh, Crack initiation under far-field cyclic compression and the

study of short fatigue cracks, 23 (1986) 953–964.

[13] T.-Y. Hsu, Z. Wang, Fatigue crack initiation at notch root under compressive cyclic

loading, Procedia Engineering, 2 (2010) 91-100.

158
[14] H. Mughrabi, The cyclic hardening and saturation behaviour of copper single crystals,

Materials Science and Engineering, 33 (1978) 207-223.

[15] F. Ackermann, L.P. Kubin, J. Lepinoux, H. Mughrabi, The dependence of dislocation

microstructure on plastic strain amplitude in cyclically strained copper single crystals, Acta

Metallurgica, 32 (1984) 715-725.

[16] J.C. Figueroa, S.P. Bhat, R. De La Veaux, S. Murzenski, C. Laird, The cyclic stress-

strain response of copper at low strains—i. Constant amplitude testing, Acta Metallurgica, 29

(1981) 1667-1678.

[17] J. Polák, M. Klesnil, Cyclic stress-strain response and dislocation structures in

polycrystalline copper, Materials Science and Engineering, 63 (1984) 189-196.

[18] V.T. Kuokkala, P. Kettunen, Cyclic stress-strain response of polycrystalline copper in

constant and variable amplitude fatigue, Acta Metallurgica, 33 (1985) 2041-2047.

[19] Z. Wang, C. Laird, Cyclic stress—strain response of polycrystalline copper under

fatigue conditions producing enhanced strain localization, Materials Science and

Engineering, 100 (1988) 57-68.

[20] J. Polák, K. Obrtlík, M. Hájek, A. Vašek, Cyclic stress-strain response of polycrystalline

copper in a wide range of plastic strain amplitudes, Materials Science & Engineering A, 151

(1992) 19-27.

[21] C. Holste, W. Kleinert, R. Gürth, K. Mecke, Cyclic stress-strain response and strain

localization effects under stress-control conditions, Materials Science & Engineering A, 187

(1994) 113-123.

[22] B. Weiss, S. Kong, R. Stickler, L. Kunz, P. Lukas, Cyclic stress-strain response and

strain localization of polycrystalline Cu tested under stress-control and different start-up

procedures, Materials Science and Engineering: A, 201 (1995) 65-76.

[23] A. Wöhler, Bericht über die versuche, welche auf der Königl. Niederschlesisch-

Märkischen eisenbahn mit apparaten zum messen der biegung und verdrehung von

159
eisenbahnwägen-achsen während der fahrt, angestellt wurden,, Zeitschrift für Bauwesen, 8

(1858) 641-652.

[24] A.S.M, Fatigue and Fracture, ASM International, USA, 1996.

[25] M.A. Meyers, K.K. Chawla, Mechanical Metallurgy: Principles and Applications,

Prentice-Hall, USA, 1984.

[26] S. Suresh, Fatigue of Metals, 2nd Edition ed., University Press, Cambridge, UK, 2003.

[27] C. Gaudin, X. Feaugas, Cyclic creep process in AISI 316L stainless steel in terms of

dislocation patterns and internal stresses, Acta Materialia, 52 (2004) 3097-3110.

[28] D. Kuhlmann-Wilsdorf, C. Laird, Dislocation behavior in fatigue, Materials Science and

Engineering, 27 (1977) 137-156.

[29] S.P. Bhat, C. Laird, The cyclic stress-strain curves in monocrystalline and

polycrystalline metals, Scripta Metallurgica, 12 (1978) 687-692.

[30] L. Llanes, J.L. Bassani, C. Laird, Cyclic response of polycrystalline copper—composite-

grain model, Acta Metallurgica Et Materialia, 42 (1994) 1279-1288.

[31] C.E. Feltner, A debris mechanism of cyclic strain hardening for F.C.C. metals,

Philosophical Magazine, 12 (1965) 1229-1248.

[32] J.M. Finney, C. Laird, Strain localization in cyclic deformation of copper single crystals,

Philosophical Magazine, 31 (1975) 339-366.

[33] A.T. Winter, A model for the fatigue of copper at low plastic strain amplitudes,

Philosophical Magazine, 30 (1974) 719-738.

[34] H. Mughrabi, Microscopic Mechanisms of Metal Fatigue, in: P. Haasen, V. Gerold, G.

Kostorz (Eds.), Pergamon, Oxford, Aachen, Federal Republic of Germany, 1980, pp. 1615.

[35] H. Mughrabi, F. Ackermann, K. Herz, Persistent slipbands in fatigued face-centred and

body-centred cubic metals, in: J.T. Fong (Ed.) Fatigue Mechanisms, American Society for

Testing and Materials 1979, pp. 69-105.

160
[36] C. Blochwitz, U. Veit, Plateau Behaviour of Fatigued FCC Single Crystals, Crystal

Research and Technology, 17 (1982) 529-551.

[37] N.Y. Jin, A.T. Winter, Cyclic deformation of copper single crystals oriented for double

slip, Acta Metallurgica, 32 (1984) 989-995.

[38] T.K. Lepistö, V.T. Kuokkala, P.O. Kettunen, Dislocation arrangements in cyclically

deformed copper single crystals, Materials Science and Engineering, 81 (1986) 457-463.

[39] B. Gong, Z. Wang, Z. Wang, Cyclic deformation behavior and dislocation structures of

[001] copper single crystals—I Cyclic stress-strain response and surface feature, Acta

Materialia, 45 (1997) 1365-1377.

[40] Z. Wang, B. Gong, Z. Wang, Cyclic deformation behavior and dislocation structures of

[001] copper single crystals—II. Characteristics of dislocation structures, Acta Materialia, 45

(1997) 1379-1391.

[41] P. Lukáš, L. Kunz, Is there a plateau in the cyclic stress-strain curves of polycrystalline

copper?, Materials Science and Engineering, 74 (1985) L1-L5.

[42] P. Lukáš, L. Kunz, Effect of grain size on the high cycle fatigue behaviour of

polycrystalline copper, Materials Science and Engineering, 85 (1987) 67-75.

[43] L. Llanes, A.D. Rollett, C. Laird, J.L. Bassani, Effect of grain size and annealing texture

on the cyclic response and the substructure evolution of polycrystalline copper, Acta

Metallurgica et Materialia, 41 (1993) 2667-2679.

[44] L. Llanes, A.D. Rollett, C. Laird, Effect of ramp-treatment on the cyclic stress-strain

curve of ―small grained‖ copper, Materials Science and Engineering: A, 167 (1993) 37-45.

[45] C.D. Liu, M.N. Bassim, D.X. You, Dislocation structures in fatigued polycrystalline

copper, Acta Metallurgica et Materialia, 42 (1994) 3695-3704.

[46] K.V. Rasmussen, O.B. Pedersen, Fatigue of copper polycrystals at low plastic strain

amplitudes, Acta Metallurgica, 28 (1980) 1467-1478.

161
[47] C.D. Liu, D.X. You, M.N. Bassim, Cyclic strain hardening in polycrystalline copper,

Acta Metallurgica et Materialia, 42 (1994) 1631-1638.

[48] P. Lukáš, M. Klesnil, Cyclic stress-strain response and fatigue life of metals in low

amplitude region, Materials Science and Engineering, 11 (1973) 345-356.

[49] P. Lukas, J. Polák, in: A.W. Thompson (Ed.) Work Hardening in Tension and Fatigue,

AIME, New York, 1977, pp. 177-205.

[50] C.E. Feltner, C. Laird, Cyclic stress-strain response of F.C.C. metals and alloys—I

Phenomenological experiments, Acta Metallurgica, 15 (1967) 1621-1632.

[51] C.E. Feltner, C. Laird, Cyclic stress-strain response of F.C.C. metals and alloys—II

Dislocation structures and mechanisms, Acta Metallurgica, 15 (1967) 1633-1653.

[52] A. Saxena, S. Antolovich, Low cycle fatigue, fatigue crack propagation and

substructures in a series of polycrystalline Cu-Al alloys, Metallurgical Transactions A, 6

(1975) 1809-1828.

[53] H. Gerber, Bestimmung der zulässigen Spannungen in Eisen-Konstructionen, Zeitschrift

des Bayerischen Architeckten und Ingenieur-Vereins 6(1874) 101-110.

[54] J. Goodman, Mechanics Applied to Engineering, Longmans Green, London, 1899.

[55] C.R. Soderberg, Factor of safety and working stress, Transactions of the American

Society of Mechanical Engineers, 52 (1939) 13-28.

[56] S.S. Manson, G.R. Halford, Practical implementation of the double linear damage rule

and damage curve approach for treating cumulative fatigue damage, International Journal of

Fracture, 17 (1981) 169-192.

[57] F. Lorenzo, C. Laird, A new approach to predicting fatigue life behavior under the

action of mean stresses, Materials Science and Engineering, 62 (1984) 205-210.

[58] R. Eckert, C. Laird, J. Bassani, Mechanism of fracture produced by fatigue cycling with

a positive mean stress in copper, Materials Science and Engineering, 91 (1987) 81-88.

162
[59] T. Wehner, A. Fatemi, Effects of mean stress on fatigue behaviour of a hardened carbon

steel, International Journal of Fatigue, 13 (1991) 241-248.

[60] Z. Xia, D. Kujawski, F. Ellyin, Effect of mean stress and ratcheting strain on fatigue life

of steel, International Journal of Fatigue, 18 (1996) 335-341.

[61] K. Kluger, T. Łagoda, Fatigue lifetime under uniaxial random loading with different

mean values according to some selected models, Materials & Design, 28 (2007) 2604-2610.

[62] K.S. Kim, H.S. Kim, C.B. Lim, Mean stress and ratcheting considerations in uniaxial

fatigue under stress control, in: X. Hu, B. Fillery, T. Qasim, K. Duann (Eds.) Structural

Integrity and Failur, 2008, Advance Materials Research, 2008, pp. 147-155.

[63] C.B. Lim, K.S. Kim, J.B. Seong, Ratcheting and fatigue behavior of a copper alloy

under uniaxial cyclic loading with mean stress, International Journal of Fatigue, 31 (2009)

501-507.

[64] D. Das, P.C. Chakraborti, Effect of stress parameters on ratcheting deformation stages of

polycrystalline OFHC copper, Fatigue & Fracture of Engineering Materials & Structures, 34

(2011) 734-742.

[65] K.F. Peters, S. Radin, A. Radin, C. Laird, Creep and fatigue interaction in

polycrystalline copper cycled under compressive mean stresses, Materials Science and

Engineering: A, 110 (1989) 115-124.

[66] N.E. Dowling, Mean stress effects in strain–life fatigue, Fatigue & Fracture of

Engineering Materials & Structures, 32 (2009) 1004-1019.

[67] P. Lukáš, L. Kunz, Effect of mean stress on cyclic stress-strain response and high cycle

fatigue life, International Journal of Fatigue, 11 (1989) 55-58.

[68] H.D. Chandler, S. Kwofie, A description of cyclic creep under conditions of axial cyclic

and mean stresses, International Journal of Fatigue, 27 (2005) 541-545.

[69] S. Kwofie, Cyclic creep of copper due to axial cyclic and tensile mean stresses,

Materials Science and Engineering: A, 427 (2006) 263-267.

163
[70] X. Yang, Low cycle fatigue and cyclic stress ratcheting failure behavior of carbon steel

45 under uniaxial cyclic loading, International Journal of Fatigue, 27 (2005) 1124-1132.

[71] G. Kang, Y. Liu, Z. Li, Experimental study on ratchetting-fatigue interaction of SS304

stainless steel in uniaxial cyclic stressing, Materials Science and Engineering: A, 435–436

(2006) 396-404.

[72] J. Pokluda, Stane, amp, x030C, P. k, Kovove Mater., 16 (1978) 583-599.

[73] V. Kliman, M. Bílý, The influence of mode control, mean value and frequency of

loading on the cyclic stress-strain curve, Materials Science and Engineering, 44 (1980) 73-

79.

[74] P. Lukáš, L. Kunz, B. Weiss, R. Stickler, W. Hessler, Effect of mean stress on the low-

amplitude cyclic stress-strain curve of polycrystalline copper, Materials Science and

Engineering: A, 118 (1989) L1-L4.

[75] P. Lukáš, L. Kunz, M. Svoboda, Stress–strain response and fatigue life of copper single

crystals cyclically loaded with a positive mean stress, Materials Science & Engineering A,

272 (1999) 31-37.

[76] F. Lorenzo, C. Laird, Strain bursts in the cyclic creep of copper single crystals at

ambient temperature, Acta Metallurgica, 32 (1984) 671-680.

[77] A.P.L. Turner, T.J. Martin, Cyclic creep of type 304 stainless steel during unbalanced

tension-compression loading at elevated temperature, Metallurgical Transactions A, 11

(1980) 475-481.

[78] R.W. Hertzberg, Deformation and Fracture Mechanics of Engineering Materials, J.

Wiley & Sons, 1996.

[79] O.H. Basquin, The exponential law of endurance tests, in, American Society of Testing

Materials Proceedings, 1910, pp. 625-630.

[80] J.D. Morrow, Internal Friction, Damping and Cyclic Plasticity, in, ASTM Special

Technical Publication, 1965, pp. 45.

164
[81] K.K. Smith, P. Watson, T.H. Topper, A stress strain function for the fatigue of metals,

Journal of Materials, JMLSA, 5 (1970) 767-778.

[82] K. Walker, Effect of environment and complex load history on fatigue life, ASTM STP

462, American Society for Testing Materials, Philadelphia, 1970.

[83] A.R. Woodward, K.W. Gunn, G. Forrest, The effect of mean stress on the fatigue of

aluminium alloys, in: International Conference on Fatigue of Metals, Institution of

Mechanical Engineers, London, 1956, pp. 158.

[84] H.C. O'Connor, J.L.M. Morrison, Effect of mean stress on the push-pull fatigue

properties of an alloy steel, in: International Conference on Fatigue of Metals, Institution of

Mechanical Engineers, London, 1956, pp. 102.

[85] G. Sines, Fatigue of materials under combined repeated stresses with superimposed

static stresses, in: Technical Note 3495, National Advisory Committee for Aeronautics,

Washington, USA, 1955.

[86] P.G. Forrest, Fatigue of Metals, Pergamon Press, Great Britain, 1962.

[87] H.O. Fuchs, R.I. Stephens, Metal Fatigue in Engineering, Wiley, New York, 1980.

[88] M. Klesnil, P. Lukas, Fatigue of Metallic Materials, Amsterdam ; New York : Elsevier

Scientific Pub. Co., 1980.

[89] W.D. Callister Jr., Materials Science and Engnieering: An Introduction, 6th ed., John

Wiley and Son, Inc, 1985.

[90] P. Lukáš, Fatigue Crack Nucleation and Microstructure, in: Fatigue and Fracture, ASM

International, USA, 1996.

[91] U. Krupp, Fatigue Crack Propagation in Metals and Alloys, Wiley-VCH, Republic of

Germany, 2007.

[92] D.K. Bullen, Steel and Its Heat Treatment, John Wiley and Sons Inc, New York, 1938.

[93] J.C. Newman Jr, E.P. Phillips, M.H. Swain, Fatigue-life prediction methodology using

small-crack theory, International Journal of Fatigue, 21 (1999) 109-119.

165
[94] T.H. Alden, W.A. Backofen, The formation of fatigue cracks in aluminum single

crystals, Acta Metallurgica, 9 (1961) 352-366.

[95] H. Mughrabi, Introduction to the viewpoint set on: Surface effects in cyclic deformation

and fatigue, Scripta Metallurgica et Materialia, 26 (1992) 1499-1504.

[96] J.Y. Mann, Fatigue of Materials:An Introductory Text, Melbourne University Press,

Australia, 1967.

[97] N. Noda, Y. Takase, Stress concentration formulae useful for any shape of notch in a

round test specimen under tension and under bending, Fatigue & Fracture of Engineering

Materials & Structures, 22 (1999) 1071-1082.

[98] P.J.E. Forsyth, Exudation of Material from Slip Bands at the Surface of Fatigued

Crystals of an Aluminium-Copper Alloy, Nature, 171 (1953) 172-173.

[99] P.J.E. Forsyth, Slip band damage and extrusion, in, Proceedings of Royal Society,

London, 1957, pp. 198-202.

[100] P.J.E. Forsyth, C.A. Stubbington, The influence of substructure on the slip observed in

pure aluminum and some aluminum alloys when subjected to fatigue stresses, Journal of the

Institute of Metals, 84 (1955) 173-175.

[101] A.H. Cottrell, D. Hull, Extrusion and intrusion by cyclic slip in copper, in, Proceedings

of Royal Society, 1957, pp. 211-213.

[102] M. Bao-Tong, C. Laird, Overview of fatigue behavior in copper single crystals—I.

Surface morphology and stage I crack initiation sites for tests at constant strain amplitude,

Acta Metallurgica, 37 (1989) 325-336.

[103] W.A. Wood, Formation of fatigue cracks, Philosophical Magazine, 3 (1958) 692-699.

[104] N.F. Mott, A theory of the origin of fatigue cracks, Acta Metallurgica, 6 (1958) 195-

197.

[105] A.J. Kennedy, Process of Creep and Fatigue in Metals, Wiley, New York, 1963.

166
[106] K. Katagiri, A. Omura, K. Koyanagi, J. Awatani, T. Shiraishi, H. Kaneshiro, Early

stage crack tip dislocation morphology in fatigued copper, Metallurgical Transactions A, 8

(1977) 1769-1773.

[107] R.E. Reed-Hill, R. Abbaschian, Physical Metallurgy Principles, 3rd ed., PWS

Publishing Company, Boston, 1994.

[108] P. Lukáš, L. Kunz, Role of persistent slip bands in fatigue, Philosophical Magazine, 84

(2004) 317-330.

[109] U. Essmann, U. Gösele, H. Mughrabi, A model of extrusions and intrusions in fatigued

metals I. Point-defect production and the growth of extrusions, Philosophical Magazine A, 44

(1981) 405-426.

[110] W.A. Wood, Conference on Fatigue, in: Bull. Inst. Met., New York: Academic Press,

Columbia, 1955.

[111] P. Neumann, Coarse slip model of fatigue, Acta Metallurgica, 17 (1969) 1219-1225.

[112] P. Neumann, Fatigue, in: Physical Metallurgy, 1983, pp. 1553-1594.

[113] M. Bao-Tong, C. Laird, Overview of fatigue behavior in copper single crystals—II.

Population, size distribution and growth kinetics of Stage I cracks for tests at constant strain

amplitude, Acta Metallurgica, 37 (1989) 337-348.

[114] J. Porter, J.C. Levy, The fatigue curves of copper, Journal of the Institute of Metals, 89

(1960) 9.

[115] W.H. Kim, C. Laird, Crack nucleation and stage I propagation in high strain fatigue—I.

Microscopic and interferometric observations, Acta Metallurgica, 26 (1978) 777-787.

[116] W.H. Kim, C. Laird, Crack nucleation and stage I propagation in high strain fatigue—

II. mechanism, Acta Metallurgica, 26 (1978) 789-799.

[117] J.C. Figueroa, C. Laird, Crack initiation mechanisms in copper polycrystals cycled

under constant strain amplitudes and in step tests, Materials Science and Engineering, 60

(1983) 45-58.

167
[118] T. Watanabe, Structural effects on grain boudary segregation, hardening and fracture,

Journal de Physique, C4 (1985) 66.

[119] N. Thompson, N. Wadsworth, N. Louat, Xi. The origin of fatigue fracture in copper,

Philosophical Magazine, 1 (1956) 113-126.

[120] R.C. Boettner, A.J. McEvily, Y.C. Liu, On the formation of fatigue cracks at twin

boundaries, Philosophical Magazine, 10 (1964) 95-106.

[121] P. Neumann, A. Tonnessen, Crack Initiation at Grain Boundaries in FCC Materials, in:

P.O. Kettunen, T.K. Lepisto, M.E. Lehtoenen (Eds.) Strength of Metals and Alloys,

Pergamon Press, Oxford, 1979, pp. 743-748.

[122] Z. Wang, H. Margolin, Mechanism for the formation of high cycle fatigue cracks at fee

annealing twin boundaries, Metallurgical Transactions A, 16 (1985) 873-880.

[123] H. Foll, High Resolution TEM, in, pp. http://www.tf.uni-

kiel.de/matwis/amat/def_en/kap_6/backbone/r6_3_4.html.

[124] T.L. Gerber, H.O. Fuchs, Analysis of non-propagating cracks in notched parts with

compressive mean stress, Journal of Materials, 3 (1968) 359-374.

[125] C.N. Reid, K. Williams, R. Hermann, Technical note: Fatigue in compression, Fatigue

of Engineering Materials and Structures, 1 (1979) 267-270.

[126] W.-Y. Chu, C.-M. Hsiao, L.-J. Jin, Fatigue crack initiation from a notch tip under a

cyclic compressive load, Scripta Metallurgica, 17 (1983) 993-996.

[127] N.A. Fleck, C.S. Shin, R.A. Smith, Fatigue crack growth under compressive loading,

Engineering Fracture Mechanics, 21 (1985) 173-185.

[128] R. Rippan, The growth of short cracks under cyclic compression, Fatigue Fracture

Engineering Material Structure, 9 (1987) 319-328.

[129] R. Rippan, The length and the shape of cracks under cyclic compression: the influence

of notch geometry, Engineering Fracture Mechanics, 31 (1988) 715-718.

168
[130] T. Zhao, Y. Jiang, Fatigue of 7075-T651 aluminum alloy, International Journal of

Fatigue, 30 (2008) 834-849.

[131] S. Suresh, Crack initiation in cyclic compression and its applications, Engineering

Fracture Mechanics, 21 (1985) 453-463.

[132] P.B. Aswath, S. Suresh, D.K. Holm, A.F. Blom, Load Interaction Effects on

Compression Fatigue Crack Growth in Ductile Solids, Journal of Engineering Materials and

Technology, 110 (1988) 278-285.

[133] D. Socie, Multiaxial fatigue damage models, Transaction of the ASME, 109 (1987)

293-298.

[134] H. Li, Q.Q. Duan, X.W. Li, Z.F. Zhang, Compressive and fatigue damage behavior of

commercially pure zinc, Materials Science and Engineering: A, 466 (2007) 38-46.

[135] J. Bauschinger, in: Mittheilunger aus dem Mechanisch-Technischen Laboratorium der

K Technischen Hochschule in Munchen, 1886, pp. 31.

[136] E. Orowan, Causes and efects of internal stresses - internal stresses and fatigue in

metals, in: Internal Stresses and Fatigue in Metals, Elsevier, Detroit, 1958.

[137] A. Abel, Historical perspectives and some of the main features of the Bauschinger

effect, Materials Forum, 10 (1987) 11-26.

[138] N.F. Mott, CXVII. A theory of work-hardening of metal crystals, Philosophical

Magazine Series 7, 43 (1952) 1151-1178.

[139] A. Seeger, Dislocation and Mechanical Properties of Crystals, John-Wiley and Sons,

New York, 1957.

[140] L.M. Brown, Orowan's explanation of the Bauschinger effect, Scripta Metallurgica, 11

(1977) 127-131.

[141] H. Ma, Bauschinger Effect and Sprinback Behaviour of Dual Phase Sheet Steels, in:

Materials Science and Engineering, University of Toronto, Toronto, 2007.

169
[142] T.M. Wu, Investigation of Bauschinger Effect in Metals, in: Department of

Mechanical Engineering, MIT, Boston, 1958.

[143] J.R. Canal, Investigation of the Bauschinger Effect in Copper, in: Department of

Mechanical Engineering, MIT, Boston, 1960.

[144] G.I. Deak, A Study on the Causes of the Bauschinger Effect, in: Department of

Mechanical Engineering, MIT, Boston, 1962.

[145] N.F. Mott, The work hardening of metals, Transactions of the Metallurgical Society of

AIME, 218 (1960) 962.

[146] P.B. Hirsch, D.H. Warrington, The flow stress of aluminum and copper at high

temperatures, Philosophical Magazine, 6 (1961) 735-767.

[147] D. Kuhlmann-Wilsdorf, A new theory of work hardening, Transactions of

Metallurgical Society of AIME, 224 (1962) 1047-1061.

[148] R.E. Smallman, Modern Physical Metallurgy, Butterworths, London, 1962.

[149] N. Ibrahim, J.D. Embury, The Bauschinger effect in single phase b.c.c. materials,

Materials Science and Engineering, 19 (1975) 147-149.

[150] E. Heyn, Metall and Erz, Heft 22 (1918) 441.

[151] G. Masing, Siemens-Koncern. Wissen Veroffent. a.d., 3 (1923).

[152] G. Masing, ibid, 5 (1926) 135.

[153] G. Masing, W. Mauksch, ibid, 5 (1926).

[154] E. Orowan, Symposium on Internal Stresses in Metals and Alloys, in, 1948.

[155] E. Schmid, W. Boas, Plasticity of Crystals, Hughes and Co., London, 1950.

[156] P. Rahifs, G. Masing, Z. F. Metallkunden, 41 (1950).

[157] N. Thompson, N.J. Wadsworth, Metal fatigue, Advances in Physics, 7 (1958) 72-169.

[158] N.H. Polakowski, E.J. Ripling, Strength and Structure of Engineering Materials,

Prentice-Hall, 1966.

170
[159] R.J. Asaro, Elastic-Plastic Memory and Kinematic-Type Hardening, Acta Metallurgica,

23 (1975) 11

[160] Z. Wang, H. Margolin, Res Mechanica, 21 (1987).

[161] ASTM Standard B170, "Standard Specification for Oxygen-Free Electrolytic Copper—

Refinery Shapes", in, ASTM International, 2010.

[162] Z. Wang, C. Laird, Dislocation structures of monocrystalline copper in cyclic latent

hardening, Metallurgical Transactions A, 20 (1989) 2033-2045.

[163] A. Radin, Z. Wang, Self-alignment grip for mechanical testing, in, Trustees of the

University of Pennsylvania, USA, 1989-07-11.

[164] Gardner's Chemical Synonyms and Trade Names, 11th Edition ed., 1999.

[165] A.S.M, Mechanical Testing and Evaluation, ASM International, USA, 2000.

[166] A.S.M, Metallography and Microstructures, ASM International, USA, 2004.

[167] S.S. Iskander, I.A.S. Mansour, G.H. Sedahmed, Electropolishing of brass alloys in

phosphoric acid, Surface Technology, 10 (1980) 357-361.

[168] L. Kunz, P. Lukáš, B. Weiss, D. Melisova, Effect of loading history on cyclic stress–

strain response, Materials Science and Engineering: A, 314 (2001) 1-6.

[169] W.H. Kim, C. Laird, The role of cyclic hardening in crack nucleation at high strain

amplitude, Materials Science and Engineering, 33 (1978) 225-231.

[170] H.J. Frost, M.F. Ashby, Deformation-mechanism maps, Pergamon Press, New York,

1982.

[171] Z.S. Basinski, S.J. Basinski, Formation and growth of subcritical fatigue cracks,

Scripta Metallurgica, 18 (1984) 851-856.

[172] Z.S. Basinski, S.J. Basinski, Low amplitude fatigue of copper single crystals—II.

Surface observations, Acta Metallurgica, 33 (1985) 1307-1317.

[173] Z.S. Basinski, S.J. Basinski, Low amplitude fatigue of copper single crystals—III. PSB

sections, Acta Metallurgica, 33 (1985) 1319-1327.

171
[174] A. Hunsche, P. Neumann, Quantitative measurement of persistent slip band profiles

and crack initiation, Acta Metallurgica, 34 (1986) 207-217.

[175] P. Neumann, The effect of surface related grain boundary stresses on fatigue, Scripta

Metallurgica et Materialia, 26 (1992) 1535-1540.

[176] F.L. Liang, C. Laird, Control of intergranular fatigue cracking by slip homogeneity in

copper I: Effect of grain size, Materials Science and Engineering: A, 117 (1989) 95-102.

[177] Z.F. Zhang, Z.G. Wang, S.X. Li, Fatigue cracking possibilty along grain boundaries

and persistent slip bands in copper bicrystals, Fatigue & Fracture of Engineering Materials &

Structures, 21 (1998) 1307-1316.

[178] Z.F. Zhang, Z.G. Wang, Grain boundary effects on cyclic deformation and fatigue

damage, Progress in Materials Science, 53 (2008) 1025-1099.

[179] T.-Y.J. Hsu, Z. Wang, Cyclic stress-strain response and microstructure evolution of

polycrystalline Cu under pure compressive cyclic loading condition, Mat. Sci. Eng. A,

(2014).

[180] Y.-D. Chuang, H. Margolin, β brass bicrystal stress-strain relations, Metallurgical

Transactions, 4 (1973) 1905-1917.

[181] H. Margolin, M. Stefan Stanescu, Polycrystalline strengthening, Acta Metallurgica, 23

(1975) 1411-1418.

[182] B.T. Ma, Z.G. Wang, A.L. Radin, C. Laird, Asymmetry behavior between tension and

compression in the cyclic deformation of copper single crystals and other ductile metals,

Materials Science & Engineering A, 129 (1990) 197-206.

172
Appendix A

Figure A 1: Strain hardening rate as a function of true strain for conditions (a) 90TCF (b)
90CTF

173
Figure A 2: Strain hardening rate as a function of true strain for conditions (a) 120TCF (b)
120CTF

174
Figure A 3: Comparison of strain hardening rate for the forward stroke between TCF and
CTF for conditions of (a) 90%σy (b) 120%σy

175

Вам также может понравиться