You are on page 1of 237

A Thesis

entitled

Influence of Surface Finish on Bending Fatigue of Forged Steel Including Heating
Method, Hardness, and Shot Cleaning Effects

by
Sean A. McKelvey

Submitted to the Graduate Faculty as partial fulfillment of the requirements for the
Masters of Science Degree in Mechanical Engineering



__________________________________________
Dr. Ali Fatemi, Committee Chair


__________________________________________
Dr. Phillip White, Committee Member


__________________________________________
Dr. Lesley Berhan, Committee Member


__________________________________________
Dr. Patricia R. Komuniecki, Dean
College of Graduate Studies



The University of Toledo

May 2011
ii















































iii
An Abstract of

Influence of Surface Finish on Bending Fatigue of Forged Steel Including Heating
Method, Hardness, and Shot Cleaning Effects

by

Sean A. McKelvey

Submitted to the Graduate Faculty as partial fulfillment of the
requirements for the Masters of Science Degree in Mechanical Engineering

The University of Toledo
May 2011


The overall objective of this study was to conduct a systematic and
comprehensive experimental investigation to evaluate and quantify forged surface finish
effect at several hardness levels (19 HRC, 25 HRC, 35 HRC, and 45 HRC) on bending
fatigue specimens of a commonly used forged steel (10B40 steel). Specimens were
subjected to reverse cantilever bending and rotating bending fatigue. Two surface
conditions were evaluated, a smooth-polished surface finish to be used as the reference
surface, and a hot-forged surface finish. The heating methods used for forging were gas
furnace heating as well as induction heating, to allow comparison of the two heating
methods, as decarburization depth differs between the two methods. Since shot blasting is
commonly used as a forged surface cleaning process with the additional benefit of
inducing compressive residual stress, the hot-forged surface finish was evaluated with
and without shot blasting. Some testing was also conducted to investigate the effect of the
flash left by the forging process. In addition, the effect of grain flow resulting from the
forging process was evaluated by testing smooth specimens machined from the same
rolled bar used for forging. Fatigue test results in this investigation confirm that the old
iv
data commonly used for the as-forged surface condition are too conservative. New forged
surface finish factors and curves as a function of hardness or tensile strength and fatigue
life were developed based on experimental data. A fracture mechanics-based approach
was also used to predict fatigue life for the as-forged fatigue specimens.



v







Acknowledgements


I would like to thank Dr. Ali Fatemi for his guidance over the last few
years. I would also like to thank the staff in the MIME machine shop, Mr. John
Jagley, Mr. Tim Grivanos and Mr. Randall Reihing. Funding for this project was
provided by the Forging Industry Educational and Research Foundation (FIERF)
and the American Iron and Steel Institute (AISI). Support and technical assistance
of Karen Lewis, George Mochenal, and Carola Sekreter of FIERF and David
Anderson of AISI are appreciated. Materials and its chemical analysis were
provided by Robert Cryderman of Gerdau-MacSteel. Induction heated forged
specimens were provided by Keystone Forging Company, courtesy of Joe
Cipriani. Specimen heat treatment and metallurgical analysis was provided by
Peter Bauerle of Chrysler Group LLC. Shot cleaning was done by Westside Sand
Blasting. Surface profiler located in the Gear Research Laboratory at OSU was
used for surface roughness measurements.






vi






Table of Contents



Abstract ..................................................................................................................... iii
Acknowledgements .................................................................................................. v
List of Tables ............................................................................................................ x
List of Figures ........................................................................................................... xiii
List of Abbreviations ............................................................................................... xxiii
List of Symbols ......................................................................................................... xxiv
1 Introduction ........................................................................................................ 1
2 Literature Review .............................................................................................. 4
2.1 Forging Process ........................................................................................... 4
2.1.1 Types of forging .............................................................................. 4
2.1.2 Effect of inclusions and grain flow in forged steel on fatigue
behavior........................................................................................... 6
2.2 Surface Finish ............................................................................................. 8
2.2.1 Decarburization ............................................................................... 9
2.2.2 Roughness measurements and parameters ...................................... 12
2.2.3 As-forged surface condition ............................................................ 14
2.2.4 Effect of surface roughness on fatigue behavior ............................. 16
2.2.5 Fatigue limit evaluation for surface finish effect ............................ 17
2.3 Sand Blasting and Shot Cleaning ................................................................ 25

vii
2.4 Heat Treatment and Hardness Effects ......................................................... 27
2.4.1 Heat treatment ................................................................................. 27
2.4.2 Hardness effects on fatigue behavior .............................................. 29
2.5 Crack Growth Behavior .............................................................................. 31
2.5.1 Fracture mechanics .......................................................................... 31
2.5.2 Fatigue crack growth........................................................................ 32
2.6 Summary ..................................................................................................... 35
3 Experimental Program, Results and Analysis................................................. 73
3.1 Material and Specimen Preparation ............................................................ 73
3.2 Testing Equipment ...................................................................................... 75
3.2.1 Monotonic tension and cantilever bending fatigue tests ................. 75
3.2.2 Rotating bending fatigue tests ......................................................... 77
3.3 Test Methods and Procedures ..................................................................... 77
3.3.1 Metallurgical analysis ..................................................................... 77
3.3.2 Surface roughness measurements ................................................... 78
3.3.3 Monotonic tension tests .................................................................. 78
3.3.4 Cantilever bending fatigue tests ...................................................... 79
3.3.5 Rotating bending fatigue tests ......................................................... 80
3.4 Experimental Results and Analysis ............................................................ 81
3.4.1 Metallurgical analysis ..................................................................... 81
3.4.2 Surface roughness characterization................................................. 82
3.4.3 Monotonic deformation behavior ................................................... 84
3.4.4 Cantilever bending fatigue behavior ............................................... 84

viii
3.4.5 Rotating bending fatigue behavior .................................................. 86
4 Comparative Analysis and Discussion ............................................................. 125
4.1 Cantilever Bending vs. Rotating Bending Fatigue ..................................... 125
4.2 Forging Flash Effect ................................................................................... 126
4.3 Effect of Hardness....................................................................................... 127
4.4 Effect of Forging Heating Method .............................................................. 129
4.5 Effect of Shot Cleaning ............................................................................... 130
4.6 Forging Grain Flow Effect .......................................................................... 131
4.7 Effect of Surface Condition ........................................................................ 132
5 Fracture Mechanics Analysis ............................................................................ 154
5.1 Crack Growth Rate Estimation ................................................................... 154
5.1.1 Crack growth behavior and rate ...................................................... 154
5.1.2 Crack growth rate versus stress intensity factor range.................... 157
5.2 Fatigue Life Prediction ............................................................................... 160
6 Prediction of Fatigue Limit and Mathematical Representation of Surface
Finish Effects ...................................................................................................... 175
6.1 Roughness-Based Prediction Models for Fatigue Limit ............................. 175
6.1.1 Murakami model ............................................................................. 175
6.1.2 Arola-Ramulu model ...................................................................... 177
6.2 Mathematical Representation of Experimental Data .................................. 179
6.2.1 Surface finish factor as a function of Brinell hardness ................... 179
6.2.2 Surface finish factor as a function of cycles to failure.................... 181

ix
6.2.3 Surface finish factor as a function of both Brinell hardness and
cycles to failure ............................................................................... 182
7 Summary and Conclusions ............................................................................... 204
References ................................................................................................................. 208
















x




List of Tables

2.1 Commonly used temperatures when heating selected ferrous alloys for hot
forging [6] ...................................................................................................... 38

2.2 Hardness variation between surface and subsurface for four types of
forged steels used in fatigue tests in [3] ......................................................... 38

2.3 Fatigue test results for four types of commonly used forged steels with the
as-forged and machined and polished surface conditions used in [3] ........... 39

2.4 Fatigue test results for a high strength forged steel from two manufacturers
with the as-forged and machined and polished surface conditions used in
[4] ................................................................................................................... 40

2.5 Relationship between surface carbon content, depth of decarburization,
and fatigue limit for a 605 M36 through-hardened steel [14] ........................ 40

2.6 Summary of endurance limits and tensile strengths for ground, machined,
hot-rolled, and as-forged surface conditions used to calculate surface
finish factors from [2] .................................................................................... 41

2.7 Values of
R
area parameter used in Murakami equation to predict fatigue
limit for specimens which have surface roughness defects [23].................... 41

2.8 Results of experimental and predicted fatigue limits using the Murakami
equation for JIS S45C steel specimens having surface roughness defects
[23] ................................................................................................................. 42

2.9 Surface finish modification factors at 10
4
and 10
6
cycles for hardened
machined surfaces calculated using the Goodman, Morrow, and SWT
models [5] ...................................................................................................... 43

2.10 Constants for use in Eqn. 2.22 for determining surface finish factors for
ground, machined, hot-rolled, and as-forged surfaces [27] ........................... 43

2.11 Fatigue test results on as-forged, as-forged/shot blasted, forged/heat
treated, and forged/ heat treated/ shot blasted specimens of EN15R steel
[1] ................................................................................................................... 43

xi

2.12 List of typical heat treatment procedures for steels [40]................................ 44

2.13 Material conditions, test parameters and fatigue lives for notched 4340
low alloy steel specimens with and without decarburization [46] ................. 45

2.14 Roughness measurements for various types of surface preparations in type
304 stainless steel [47] ................................................................................... 45

3.1 Material composition of 10B40 steel (Courtesy of Robert Cryderman of
Gerdau-MacSteel) .......................................................................................... 88

3.2 Summary of surface roughness measurements (m) of as-forged, as-
forged and heat treated, and as-forged and heat treated and shot cleaned
for 10B40 steel ............................................................................................... 88

3.3 Summary of monotonic tension test results ................................................... 89

3.4 Summary of fatigue test results with different heating methods and surface
finish conditions for 45 HRC specimens ....................................................... 90

3.5 Summary of fatigue test results with different heating methods and surface
finish conditions for 35 HRC specimens ....................................................... 91

3.6 Summary of fatigue test results with different heating methods and surface
finish conditions for 25 HRC specimens ....................................................... 92

3.7 Summary of fatigue test results with different heating methods and surface
finish conditions for 19 HRC specimens ....................................................... 93

3.8 S-N line equation constants and fatigue strengths ......................................... 94

4.1 Summary of calculated surface finish factors for as-forged gas furnace
heated, as-forged induction heated, as-forged and shot cleaned gas furnace
heated, and as-forged and shot cleaned induction heated specimens, as
well as historical surface finish factors .......................................................... 137

5.1 Summary of threshold stress intensity factor range and transition crack
depth calculations for as-forged specimens forged at different hardness
levels .............................................................................................................. 163

6.1 Summary of calculations of square root area parameters and fatigue limit
predictions using Murakami model for as-forged specimens ........................ 187


xii
6.2 Summary of calculations of stress concentration factors, fatigue notch
factors, and fatigue limit predictions using Arola-Ramulu and Neuber
models for as-forged specimens ..................................................................... 188

6.3 Summary of R
2
values used in three dimensional fits to the surface finish
factor data as a function of hardness and cycles to failure for the different
surface conditions tested ................................................................................ 189

6.4 Summary of constants for use in Eqn. 6.17 for the three dimensional
surface finish data fits shown in Figures 6.9 through 6.12 ............................ 189




































xiii







List of Figures

2-1 Illustration of (a) closed-die [7] and (b) open-die [8] showing how
material protrudes between upper and lower die ........................................... 46

2-2 The relative shape change of treated and untreated inclusions compared to
shape change of the steel billet for strains of 0.2 and 1.0 at 1000 C

[11] .... 47

2-3 Illustration of a cross section of a forged steel part showing grain flow
[12] .............................................................................................................47

2-4 Projection of a defect on to a plane perpendicular to maximum tensile
stress which is used to calculate the square root area parameter [22] ........... 48

2-5 Endurance limit versus tensile strength for as-forged surface finish and
polished surface finish specimens [36] .......................................................... 48

2-6 Brinell hardness versus distance from surface for as-forged specimens
with average core hardness of (a) 185 HB, (b) 200 HB, (c) 260 HB, and
(d) 360 HB [2] ................................................................................................ 50

2-7 Magnified images (150 x magnification) of a cross section of as-forged
specimens showing surface decarburization with average core hardness of
(a) 185 HB and decarburization ranging from 0.64 mm to 0.76 mm, (b)
200 HB and decarburization ranging from 0.25 mm to 0.30 mm, (c) 260
HB and decarburization ranging from 0.76 mm to 0.89 mm, and (d) 360
HB and decarburization ranging from 0.30 mm to 0.38 mm [2] ................... 51

2-8 Endurance limit versus tensile strength showing upper and lower limits for
as-forged surface specimens [13]................................................................... 52

2-9 Definition of (a) R
a
, (b) R
z
, (c) R
t
and R
y
, and (d) S
m
to characterize surface
roughness [18] ................................................................................................ 53

2-10 Maximum working stress (at 10
7
cycles) versus mean working stress (at
10
7
cycles) for the ground, machined, hot-rolled and as-forged surface
conditions of steel (302 HB-321 HB) [2] ...................................................... 54

2-11 Relationship between Brinell hardness and tensile strength for steels [2]..... 54

xiv

2-12 Plot of endurance limit (defined at 10
7
cycles) versus tensile strength for
steels having ground, machined, hot-rolled and as-forged surface
conditions of steels [2] ................................................................................... 55

2-13 Surface roughness versus median fatigue life for SAE 1035 and SAE 3130
steels [20] ....................................................................................................... 55

2-14 Scatter of fatigue test results for SAE 1035 and SAE 3130 steels having a
range of surface roughness [20] ..................................................................... 56

2-15 Configuration and dimensions (mm) of JIS S45C steel fatigue specimens
used in [23] .................................................................................................... 57

2-16 Magnified image of artificially induced surface roughness for annealed
(A) and quenched and tempered (QT) JIS S45C steel with roughness pitch
at 100, 150 and 200 in [23] ............................................................................ 57

2-17 Composite plot of stress versus fatigue life for different types of roughness
in JIS S45C steel specimens [23] ................................................................... 58

2-18 Relationship between
R
area /2b and a/2b for annealed (A) and
quenched and tempered (QT) JIS S45C steel fatigue specimens [23] ........... 58

2-19 Relationship between predicted and experimental fatigue limits for
annealed and quenched and tempered JIS S45C steel [23] ............................ 59

2-20 Specimen geometry and test set up for the hardened SAE 4140 steel shaft
subjected to bending fatigue in [5]................................................................. 59

2-21 Surface finish modification factors for fatigue life ranging from 10
4
to 10
6

cycles for hardened machined surfaces calculated using the Goodman,
Morrow, and SWT models for hardened machined steel surfaces [5] ........... 60

2-22 Roughness profile showing definition of the parameter (average radius
of the dominant profile valleys) used in Arola-Ramulu Model [25] ............. 60

2-23 Fatigue stress concentration factor calculated using Arola-Ramulu and
Neuber models, and experimental fatigue stress concentration factor
versus average surface roughness for a high strength low alloy steel [25] .... 61

2-24 Surface finish modification factor for steels versus tensile strength or
Brinell hardness for ground and polished, machined, hot-rolled, and as-
forged surface conditions [26] ....................................................................... 61


xv
2-25 Stress versus cycles to failure for two type of as-forged surface finish
specimens made from SAE 1035 steel (233 HB-280 HB) and SAE 4063
steel (388 HB-444 HB) [13] .......................................................................... 62

2-26 S-N curves for steel specimens having machined and polished surface and
as-forged surface conditions .......................................................................... 62

2-27 Atomic structure of iron showing (a) body centered cubic, and (b) face
centered cubic [43] ......................................................................................... 63

2-28 Phase diagram for plain carbon steels [43] .................................................... 63

2-29 Hardness versus carbon concentration in plain carbon steels [43] ................ 64

2-30 Tensile strength, yield strength, and reduction in area versus tempering
temperature for 4340 oil quenched steel with martensitic microstructure
[43] ................................................................................................................. 65

2-31 Monotonic and cyclic stress-strain curves for (a) SAE 1045 steel at 595
HB, 500 HB, 450 HB, and 390 HB, as well as (b) SAE 4142 steel at 670
HB, 560 HB, 475 HB, 450 HB, and 380 HB [44] ......................................... 66

2-32 Stress amplitude versus plastic strain amplitude for SAE 1045 steel at 595
HB, 500 HB, 450 HB, and 390 HB, and for SAE 4142 steel at 670 HB,
560 HB, 475 HB, 450 HB, and 380 HB [44] ................................................. 67

2-33 Total strain amplitude versus cycles to failure for strain-controlled
completely-reversed fatigue testing of (a) SAE 1045 steel at 700 HB, 600
HB, and 450 HB, as well as (b) SAE 4142 steel at 670 HB, 560 HB, and
450 HB [44] ................................................................................................... 67

2-34 Total strain amplitude versus hardness for different cycles to failure for
strain-controlled completely-reversed fatigue testing of (a) SAE 1045
steel, and (b) SAE 4142 steel [44] ................................................................. 68

2-35 Plot of growth rate (da/dN) versus stress intensity factor range (K) for
martensitic steels having various yield and tensile strengths [45] ................. 69

2-36 Specimen configuration and dimensions for a 4340 low alloy steel notched
specimen used to determine the effect of decarburization on fatigue
behavior in [46] .............................................................................................. 69

2-37 Magnified image of surface roughness for type 304 stainless steel having
(a) ground and polished surface, and surface roughened with (b) 600 grit,
(c) 240 grit, and (d) 50 grit sand paper [47] ................................................... 70


xvi
2-38 Crack length versus number of cycles for type 304 stainless steel
specimens at four levels of surface roughness [47] ....................................... 71

2-39 Crack initiation life versus surface roughness for type 304 stainless steel
specimens [47] ............................................................................................... 71

2-40 Notch geometry of three-point bending steel specimen with hardness
between 180 HB and 230 HB used for crack growth study in [48] ............... 72

2-41 Number of cycles versus average roughness for three-point bending steel
specimens with hardness between 180 HB and 230 HB [48] ........................ 72

3-1 Photograph of (a) gas furnace heated specimen, (b) induction heated
specimen, and (c) machined and polished specimen ..................................... 95

3-2 Specimen configuration and nominal dimensions (mm) for as-forged
surface finish specimens subjected to (a) cantilever bending fatigue, and
(b) rotating bending fatigue ........................................................................... 96

3-3 Flow charts showing (a) specimen preparation, and various testing
conditions for (b) specimens machined form rolled bar, (c) gas furnace
heated forged specimens, and (d) induction heated forged specimens .......... 99

3-4 Test setup used for reversed cantilever bending fatigue testing of as-
forged, shot-cleaned, and machined and polished specimens........................ 100

3-5 Plot of calculated strain versus measured strain used to verify cantilever
bending test setup ........................................................................................... 101

3-6 Four-point rotating bending fatigue testing machine ..................................... 102

3-7 Cross section of specimen gage section area prior to heat treatment
showing a mixture of ferrite and pearlite (Courtesy of Peter Bauerle of
Chrysler) for (a) induction heated specimens, and (b) gas furnace heated
specimens ....................................................................................................... 103

3-8 Magnification of gage section area showing decarburization for gas
furnace heated specimens with martensitic microstructure (Courtesy of
Peter Bauerle of Chrysler) at (a) 45 HRC, (b) 35 HRC, (c) 25 HRC, and
(d) 19 HRC ..................................................................................................... 104

3-9 Magnification of gage section area showing decarburization for induction
heated specimens at 45 HRC with martensitic microstructure (Courtesy of
Peter Bauerle of Chrysler) ............................................................................. 104


xvii
3-10 Grain flow resulting from forging process for induction heated specimens
(Courtesy of Peter Bauerle of Chrysler) ........................................................ 105

3-11 Magnified image showing surface irregularities for induction heated
specimens prior to heat treatment (Courtesy of Peter Bauerle of Chrysler) .. 105

3-12 Brinell hardness versus depth below the surface for induction heated
forged and gas furnace heated forged specimens .......................................... 106

3-13 Roughness profiles for induction heated specimens (a) prior to heat
treatment, (b) as-forged at 45 HRC, and (c) as-forged and shot cleaned at
45 HRC .......................................................................................................... 107

3-14 Roughness profiles for gas furnace heated specimens (a) prior to heat
treatment, (b) as-forged at 45 HRC, and (c) as-forged and shot cleaned at
45 HRC .......................................................................................................... 108

3-15 Roughness versus hardness for shot cleaned and as-forged surfaces for (a)
roughness parameter R
a
, (b) roughness parameter R
t
, (c) roughness
parameter R
y
, and (d) roughness parameter R
z
............................................... 109

3-16 Monotonic stress-strain curves at different hardness levels for 10B40 steel . 110

3-17 Tensile strength versus Brinell hardness for 10B40 steel .............................. 110

3-18 Typical fracture surface of as-forged specimens subjected to fully-reversed
cantilever bending fatigue at (a) stress amplitude of 927 MPa at 45 HRC,
(b) stress amplitude of 301 MPa at 45 HRC, (c) stress amplitude of 610
MPa at 35 HRC, (d) stress amplitude of 251 MPa at 35 HRC, (e) stress
amplitude of 609 MPa at 25 HRC, (f) stress amplitude of 260 MPa at 25
HRC, (h) stress amplitude of 500 MPa at 19 HRC, and (i) stress amplitude
of 249 MPa at 19 HRC................................................................................... 111

3-19 Displacement amplitude versus normalized cycles in cantilever bending
fatigue tests for 45 HRC specimens (a) as-forged with induction heat, (b)
as-forged with gas furnace heat, (c) as-forged with induction heat and
shot-cleaned, and (d) as-forged with gas furnace heat and shot cleaned ....... 112

3-20 Displacement amplitude versus normalized cycles in cantilever bending
fatigue tests for 35 HRC specimens (a) as-forged with induction heat, (b)
as-forged with gas furnace heat, (c) as-forged with induction heat and
shot-cleaned, and (d) as-forged with gas furnace heat and shot cleaned ....... 113

3-21 Displacement amplitude versus normalized cycles in cantilever bending
fatigue tests for 25 HRC specimens (a) as-forged with induction heat, (b)

xviii
as-forged with gas furnace heat, (c) as-forged with induction heat and
shot-cleaned, and (d) as-forged with gas furnace heat and shot cleaned ....... 114

3-22 Displacement amplitude versus normalized cycles in cantilever bending
fatigue tests for 19 HRC specimens (a) as-forged with induction heat, (b)
as-forged with gas furnace heat, (c) as-forged with induction heat and
shot-cleaned, and (d) as-forged with gas furnace heat and shot cleaned ....... 115

3-23 Displacement amplitude versus normalized cycles in cantilever bending
fatigue tests for machined and polished specimens at (a) 45 HRC, (b) 35
HRC, (c) 25 HRC, and (d) 19 HRC ............................................................... 116

3-24 Stress-life curves under cantilever bending fatigue for as-forged induction
heated specimens at (a) 45 HRC, (b) 35 HRC, (c) 25 HRC, and (d) 19
HRC ............................................................................................................... 117

3-25 Stress-life curves under cantilever bending fatigue for as-forged and shot
cleaned induction heated specimens at (a) 45 HRC, (b) 35 HRC, (c) 25
HRC, and (d) 19 HRC .................................................................................... 118

3-26 Stress-life curves under cantilever bending fatigue for as-forged gas
furnace heated specimens at (a) 45 HRC, (b) 35 HRC, (c) 25 HRC, and (d)
19 HRC .......................................................................................................... 119

3-27 Stress-life curves under cantilever bending fatigue for as-forged and shot
cleaned gas furnace heated specimens at (a) 45 HRC, (b) 35 HRC, (c) 25
HRC, and (d) 19 HRC .................................................................................... 120

3-28 Stress-life curves under cantilever bending fatigue for machined and
polished specimens at (a) 45 HRC, (b) 35 HRC, (c) 25 HRC, and (d) 19
HRC ............................................................................................................... 121

3-29 Midlife load versus displacement curves for highest load amplitude tests at
45 HRC, 35 HRC, 25 HRC, and 19 HRC for (a) induction heated forged
specimens, and (b) machined and polished specimens .................................. 122

3-30 Typical fracture surface of as-forged specimens subjected to rotating
bending fatigue at (a) stress amplitude of 551 MPa at 35 HRC, (b) stress
amplitude of 229 MPa at 35 HRC, (c) stress amplitude of 502 MPa at 19
HRC, and (d) stress amplitude of 227 MPa at 19 HRC ................................. 123

3-31 Stress-life curves under rotating bending fatigue for (a) gas furnace heated
as-forged surface at 35 HRC, (b) machined and polished surface at 35
HRC, (c) gas furnace heated as-forged surface at 19 HRC, and (d)
machined and polished surface at 19 HRC .................................................... 124


xix
4-1 Cantilever bending versus rotating bending for as-forged and polished
surface conditions at (a) 35 HRC, and (b) 19 HRC ....................................... 138

4-2 Forging flash effect on fatigue behavior of as-forged specimens with gas
furnace heating in cantilever bending tests for (a) 45 HRC, and (b) 25
HRC ............................................................................................................... 139

4-3 Effect of hardness on fatigue behavior of machined and polished
specimens in cantilever bending .................................................................... 140

4-4 Effect of hardness on cantilever bending fatigue behavior for as-forged
specimens with (a) induction heating, and (b) gas furnace heating ............... 141

4-5 Effect of hardness on cantilever bending fatigue behavior for shot cleaned
forged specimens with (a) induction heating, and (b) gas furnace heating ... 142

4-6 Effect of heating method on cantilever bending fatigue for as-forged
surface specimens at (a) 45 HRC, (b) 35 HRC, (c) 25 HRC, and (d) 19
HRC ............................................................................................................... 143

4-7 Effect of heating method on cantilever bending fatigue for shot cleaned
forged surface specimens at (a) 45 HRC, (b) 35 HRC, (c) 25 HRC, and (d)
19 HRC .......................................................................................................... 144

4-8 Effect of shot cleaning on cantilever bending fatigue for induction heated
forged specimens at (a) 45 HRC, (b) 35 HRC, (c) 25 HRC, and (d) 19
HRC ............................................................................................................... 145

4-9 Effect of shot cleaning on cantilever bending fatigue for gas furnace
heated forged specimens at (a) 45 HRC, (b) 35 HRC, (c) 25 HRC, and (d)
19 HRC .......................................................................................................... 146

4-10 Comparison of cantilever bending fatigue behavior between forged
specimens with forged surface removed (circular symbols) and machined
specimens from rolled bars (triangular symbols) at 45 HRC and 25 HRC
hardness levels ............................................................................................... 147

4-11 Surface finish effect on cantilever bending fatigue behavior at 45 HRC
hardness level with induction and gas furnace heating for (a) as-forged,
and (b) as-forged and shot cleaned conditions ............................................... 148

4-12 Surface finish effect on cantilever bending fatigue behavior at 35 HRC
hardness level with induction and gas furnace heating for (a) as-forged,
and (b) as-forged and shot cleaned conditions ............................................... 149


xx
4-13 Surface finish effect on cantilever bending fatigue behavior at 25 HRC
hardness level with induction and gas furnace heating for (a) as-forged,
and (b) as-forged and shot cleaned conditions ............................................... 150

4-14 Surface finish effect on cantilever bending fatigue behavior at 19 HRC
hardness level with induction and gas furnace heating for (a) as-forged,
and (b) as-forged and shot cleaned conditions ............................................... 151

4-15 Composite plot of surface finish factor at 10
6
cycles versus tensile strength
or hardness showing differences between induction and gas furnace
heating forging and old and current data with (a) as-forged surface, and (b)
as-forged and shot cleaned surface ................................................................ 152

4-16 Predicted fatigue life based on historical data (cycles) versus experimental
fatigue life (cycles) for as-forged, and as-forged and shot cleaned
specimens at 45 HRC, 35 HRC, 25 HRC, and 19 HRC hardness levels ....... 153

5-1 Plots of (a) measured crack length and (b) calculated crack depth versus
change in displacement amplitude for as-forged cantilever bending
specimens at 45 HRC, 35 HRC, 25 HRC and 19 HRC ................................. 164

5-2 Measured crack depth versus measured crack length at fracture for 45
HRC, 35 HRC, 25 HRC and 19 HRC specimens having an as-forged
surface from cantilever bending fatigue tests ................................................ 165

5-3 Schematic illustrations for as-forged surface specimens plotted in Figure
5.2 of (a) crack depth, a, and crack length, b, and (b) crack shape ................ 165

5-4 Change in displacement amplitude versus normalized cycles in cantilever
bending fatigue tests of as-forged surface specimens at (a) 45 HRC
hardness level, (b) 35 HRC hardness level, (c) 25 HRC hardness level, and
(d) 19 HRC hardness level ............................................................................. 166

5-5 Calculated crack depth (using Eqn. 5.3) versus normalized cycles in
cantilever bending fatigue tests of as-forged surface specimens at (a) 45
HRC hardness level, (b) 35 HRC hardness level, (c) 25 HRC hardness
level, and (d) 19 HRC hardness level ............................................................ 167

5-6 Schematic plot of crack growth rate versus stress intensity factor range
[26] ................................................................................................................. 168

5-7 Crack growth rate (da/dN) versus stress intensity factor range (K) for
forged specimens subjected to cantilever bending fatigue at (a) 45 HRC,
(b) 35 HRC, (c) 25 HRC, and (d) 19 HRC .................................................... 169


xxi
5-8 Composite plot of crack growth rate (da/dN) versus stress intensity factor
range (K) for 45 HRC, 35 HRC, 25 HRC, and 19 HRC forged specimens
subjected to cantilever bending fatigue .......................................................... 170

5-9 Schematic plot of a Kitagawa-Takahashi diagram giving the relationship
between stress range and crack length [26] ................................................... 171

5-10 Schematic plot of crack growth rate (da/dN) versus stress intensity factor
range (K) showing effect of short cracks and extrapolation of Paris
equation line [26] ........................................................................................... 171

5-11 Estimated cycles to failure versus experimental cycles to failure for as-
forged at 45 HRC, 35 HRC, 25 HRC and 19 HRC specimens using
average A and n values from [46] in Paris equation ...................................... 172

5-12 Estimated cycles to failure versus experimental cycles to failure for as-
forged specimens using experimental A and n values at (a) 45 HRC and 35
HRC, and (b) 25 HRC and 19 HRC ............................................................... 173

5-13 Plot of estimated cycles to failure versus experimental cycles to failure
for 45 HRC forged specimens showing effect of assumed K
c
value in life
predictions ...................................................................................................... 174

6-1 Predicted fatigue limit versus experimental fatigue limit using Murakami
model for as-forged specimens at 45 HRC, 35 HRC, 25 HRC, and 19 HRC 190

6-2 Predicted fatigue limit versus experimental fatigue limit for as-forged
specimens at 45 HRC, 35 HRC, 25 HRC, and 19 HRC using (a) Arola-
Ramulu model, and (b) Neuber model ........................................................... 191

6-3 Surface finish factor (k
s
) versus hardness at different fatigue lives for as-
forged specimens with (a) induction heating, and (b) gas furnace heating ... 192

6-4 Surface finish factor (k
s
) versus hardness at different fatigue lives for shot
cleaned specimens with (a) induction heating, and (b) gas furnace heating .. 193

6-5 Surface finish factor versus Brinell hardness for as-forged and as-forged
and shot cleaned specimens with gas furnace and induction heating at 10
6
,
3x10
5
, 10
5
, and 3x10
4
cycles to failure .......................................................... 194

6-6 Surface finish factor (k
s
) versus cycles to failure at different hardness
levels for as-forged specimens with (a) induction heating, and (b) gas
furnace heating ............................................................................................... 195


xxii
6-7 Surface finish factor (k
s
) versus cycles to failure at different hardness
levels for shot cleaned specimens with (a) induction heating, and (b) gas
furnace heating ............................................................................................... 196

6-8 Surface finish factor versus cycles to failure for as-forged and as-forged
and shot cleaned specimens with gas furnace and induction heating at 45
HRC (421 HB), 35 HRC (327 HB), 25 HRC (253 HB), and 19 HRC (220
HB) ................................................................................................................. 197

6-9 3D plot showing surface finish factor as a function of hardness and cycles
to failure for (a) induction heated forged, and (b) gas furnace heated
forged specimens ........................................................................................... 198

6-10 3D plot showing surface finish factor as a function of hardness and cycles
to failure for (a) induction heated forged and shot cleaned, and (b) gas
furnace heated forged and shot cleaned specimens ....................................... 199

6-11 3D plot showing surface finish factor as a function of hardness and cycles
to failure for induction heated and gas furnace heated forged specimens
with (a) as-forged surface, and (b) as-forged and shot cleaned specimens ... 200

6-12 3D plot showing surface finish factor as a function of hardness and cycles
to failure for induction heated and gas furnace heated forged specimens
with as-forged surface and as-forged and shot cleaned surface ..................... 201

6-13 Surface finish factor from experimental data and old data versus surface
finish factor calculated from fit to experimental data (both induction and
gas furnace heated specimens) for failure ranging from 10
4
to 10
6
cycles
for (a) as-forged surface, and (b) as-forged and shot cleaned surface.
Scatter bands of 5% and 10% are shown ....................................................... 202

6-14 Surface finish factor versus hardness for 10
4
, 10
5
, and 10
6
cycles to failure
for (a) as-forged surface using Eqn. 6.18, and (b) as-forged and shot
cleaned surface using Eqn. 6.19 ..................................................................... 203



xxiii







List of Abbreviations


EPFM elastic-plastic fracture mechanics
HB Brinell hardness number
HCF high cycle fatigue
HRB, HRC Rockwell B-scale, C-scale number
HV Vickers hardness number
LCF low cycle fatigue
LEFM linear elastic fracture mechanics
RMS root mean square


























xxiv







List of Symbols


geometry factor used in calculation of stress intensity factor
a material constant used to determine the notch sensitivity
displacement amplitude

a
change in displacement amplitude
true strain range
K stress intensity factor range
K
th
threshold stress intensity factor

e
true elastic strain

f
fatigue ductility coefficient

p
true plastic strain
ratio between spacing and height of surface irregularities
notch radius
average notch radius determined from the dominate roughness profile
valleys
true stress

a
stress amplitude

f
fatigue strength coefficient

w
fatigue limit predicted in area parameter model

w
,
wo
fatigue limit, fatigue strength

a crack length or depth
a
0
initial crack length
a
d
length of decarburized layer
a
cr
critical crack length

xxv
A fatigue strength coefficient or constant in Paris equation for intercept of line
B fatigue strength exponent
d diameter
d
e
effective diameter used for nonrotating bending size factor
e engineering strain
E modulus of elasticity
K monotonic strength coefficient
k
a
, k
s
surface finish modification factor
k
b
size factor
k
c
loading factor
f
K effective fatigue notch factor in Arola-Ramulu equation
t
K effective stress concentration factor in Arola-Ramulu equation
M
a
alternating moment
n strain hardening exponent, slope in Paris equation, or loading factor in
Arola-Ramulu equation
N number of cycles
N
b
cycles to grow crack to critical length
N
d
cycles to grow crack through decarburized layer
N
f,
N
t
cycles to failure
q notch sensitivity
R stress ratio
R
a
arithmetic average height roughness parameter
R
p
maximum peak height roughness parameter
R
q
root mean square roughness parameter
R
t
maximum height of profile roughness parameter
R
v
maximum depth of valley roughness parameter
R
y
largest peak to valley height roughness parameter
R
z
the ten-point height roughness parameter
S engineering stress
S
a
alternating stress
S
e
, S
f
fatigue limit

xxvi
S
m
mean stress or mean spacing between profile peaks at the mean line
S
u
ultimate tensile strength
S
y
yield strength






1








Chapter 1


Introduction






Fatigue fractures are the most common type of mechanical failures of components
and structures. Fatigue failures typically initiate at the surface where micro-cracks form.
A majority of life consists of crack initiation in the case of a smooth and polished surface.
Depending on the surface finish, micro-cracks may already be present, resulting in a
significant reduction of fatigue life.
Forgings are typically accompanied by surface roughness and decarburization. It
has been shown by many investigators that as surface roughness increases there is a
significant decrease in fatigue life in the high cycle fatigue (HCF) region. Surface
roughness has less of an effect in the low cycle fatigue (LCF) region due to a high degree
of plastic deformation. Decarburization reduces the hardness in the decarburized area.
The reduction in hardness results in a reduction of strength, but an increase in ductility.
Decarburization normally results in reduced fatigue life in the HCF region. However, the
reduction is dependent upon the depth of the decarburization.
The correction factors typically used in fatigue design and implemented in many
mechanical design textbooks to correct for the as-forged surface condition are typically
based on data published in the 1940s. It has been found by several investigators that the

2
existing data for as-forged surface condition is too conservative. Such conservative
values often result in over-engineered designs of many forged parts, leading not only to
increased cost, but also inefficiencies associated with increased weight, such as increased
fuel consumption in the automotive industry. In addition, this can reduce forging
competitiveness as a manufacturing process in terms of cost and performance prediction
in the early design stage, compared to alternative manufacturing processes. Although
many forgings are machined in critical areas to remove the forged surface, some forgings
are not machined in all critical areas subsequent to forging, such as the shank areas in
forged connecting rods. Therefore, the overall objective of this project was to conduct a
systematic and comprehensive study to evaluate and quantify forged surface finish effect,
with and without residual stresses, at several hardness levels (19 HRC, 25 HRC, 35 HRC,
and 45 HRC), on bending fatigue specimens of a commonly used forged steel.
Specimens were subjected to reverse cantilever bending and rotating bending
fatigue. Two surface conditions were evaluated; a smooth-polished surface finish to be
used as the reference surface, and a hot-forged surface finish, as hot forging is the most
common type of forging. The heating methods used for forging were gas furnace heating
as well as induction heating, both common heating methods, to allow comparison of the
two, as surface roughness and decarburization depths differ between the two methods.
Shot blasting is commonly used as a surface cleaning process of forgings with the
additional benefit of inducing surface compressive residual stress. Since residual stresses
play a critical synergistic role with surface finish, the hot-forged surface finish is
evaluated with and without residual compressive stresses from shot cleaning. Some
testing was also conducted to investigate the effect of the flash left by the forging

3
process. This was done by placing the flash at the location of maximum stress under
cantilever bending conditions. However, in the majority of the tests, the flash line was
located at the neutral axis. In addition, the effect of grain flow resulting from the forging
process was evaluated by testing smooth specimens machined from the same rolled bars
used for forging of the specimens.
Chapter 2 presents a literature review on surface finish effects on forgings. This
includes the effect of decarburization, hardness, and surface roughness on fatigue
behavior of steels. Chapter 3 describes the experimental procedure used and results
obtained. Specimen preparation, metallurgical analysis, roughness measurements, testing
equipment and procedures are all discussed. Chapter 4 covers a comparative analysis and
discussion of results. Comparisons of loading method and heating method used are made.
The effects of forging flash line, grain flow, hardness and surface finish are also
discussed. In addition, the surface finish factors calculated in the current investigation are
compared to the surface finish factors based on old data currently available for design.
Chapter 5 describes a fracture mechanics analysis of the fatigue data obtained in this
study. In this analysis, crack growth rates and fatigue lives of the as-forged surface
specimens are predicted. Chapter 6 discusses data fits and mathematical representation of
the data. Chapter 6 also contains fatigue limits calculated from exiting prediction models
discussed in the literature review and compares the calculations to fatigue data. Finally,
Chapter 7 summarizes the findings from this investigation.







4







Chapter 2





Literature Review




Forging is a commonly used manufacturing method. Vehicles are made of up to
25% forged components [1]. The existing information for the as-forged surface condition
effect on fatigue behavior is too conservative. Although significant improvements to the
steel making and forging processes have been made, the correction factors used in fatigue
design for the as-forged surface condition are still based on data published in the 1930s
and 1940s [2-4]. This results in an over-engineered design of many mechanical parts,
which is accompanied by an increase in production cost [5]. This chapter reviews
previous studies on the effects of surface condition, most importantly the as-forged
surface condition, on the fatigue behavior of steels.

2.1 Forging Process
2.1.1 Types of forging
Forging is a manufacturing process where compressive stresses are used to form
or shape material. There are three types of forging; cold forging, warm forging, and hot
forging [6]. Cold, warm, and hot refer to the temperatures at which the process takes
place. Cold forging typically takes place around room temperature. Hot forging is done at

5
a temperature which allows recrystallization of the material being forged, typically
around 1200C for steels. Forging above the recrystallization temperature requires less
power due to increased ductility. Most forgings are hot forged due to the fact that it
allows for a more complex geometry than cold or warm forging. In addition, cold forging
can limit the size of the part being forged. Examples of parts hot forged include crank
shafts, connecting rods, gears, and tools. When hot forging, the material can be heated up
by use of gas furnace or induction heating. Typically induction heating is thought to
produce a higher quality forging than gas furnace heating. The specimens used in the
current investigation were hot forged by use of both induction heat and gas furnace heat.
Induction heating is where electrical currents are used to heat up the material, and
only the material, unlike gas furnace heating [7]. When heating by gas furnace, the whole
atmosphere inside the furnace needs to be brought to temperature in order to heat the
material. Gas furnace heating is a long process when compared to induction heating,
which heats the material more rapidly and uniformly. In addition, the gas furnace can
produce a poorer quality of surface finish than induction heating; more scale formation,
oxidation, decarburization and grain coarsening [7]. The decarburization and scale
formation is significantly decreased for induction heating. The gas furnace also takes up
more room and produces more negative byproducts than an induction heater. Therefore,
induction heating is generally a more consistent and efficient process than gas furnace
heating.
Hot forging begins with a billet or a bar. As the material is heated up its yield
strength is decreased and ductility is increased, making it easier to form. Table 2.1 lists
typical forging temperatures for some steels [7]. Flow stress is an important factor in

6
determining how a material can be formed [8]. Flow stress, or yield stress, is a function
of temperature, strain, strain rate, and microstructure of the billet or bar. Since hot forging
is done at temperatures above the recrystallization temperature the flow stress is mostly
affected by temperature and strain rate, and not by strain, thus reducing strain hardening.
Closed-die forging is done by placing a heated billet in between an upper and
lower die and then bringing the upper and lower dies together. The heated material then
fills the cavity in the die. In many cases excess material is allowed to protrude from a
narrow gap between the upper and lower dies creating the flash (see Figure 2-1(a)). It is
possible to produce closed-die forgings without flash; however this is a much more
controlled, therefore more costly process and may not be necessary for most applications.
Open-die forging is a similar process involving upper and lower dies. However, the dies
are typically flat or have a simple contour. In addition, material expansion is not
restricted in the plane perpendicular to the loading direction (see Figure 2-1(b)) [9].
Open-die forgings also require larger tolerances compared to closed-die forgings and
require more machining for close tolerances [10]. It should be noted that the specimens in
the current study were forged using a closed-die forging process.

2.1.2 Effect of inclusions and grain flow in forged steel on fatigue behavior
In a study by Collins and Michal [11] tension-tension fatigue tests (R = 0.1) were
performed on forged specimens made of AISI 4140 steel in order to determine the effects
of changes in shape and distribution of MnS inclusions. Three types of AISI 4140 steel
were used, IGS/HS, IGS, and IGS/SC, each with a different composition. The specimens
were machined from a forged block of material in both longitudinal and transverse

7
directions. The specimens were then quenched and tempered to a hardness of 38 HRC
with a tempered martensite microstructure. Finally, they were polished to a 0.25 m
surface finish. It should be noted that the specimens had compressive residual stresses at
the surface around 357 MPa resulting from machining [11]. During the forging process, it
is possible for the inclusions to be reoriented following flow lines. It is also possible for
the inclusions to deform more than the material surrounding it during forging. They
describe a method of controlling the shape of an inclusion, which has been shown to
improve the fatigue life. Different chemicals and thermal treatments can be used. Adding
calcium can harden the inclusions to prevent deformation during working. Figure 2-2
shows the shape change of MnS inclusions and Ca treated MnS inclusions. It can be seen
that the treated inclusions experienced little change in shape. The results from fatigue
testing showed that forging improved the fatigue behavior of the base material. The Ca
treatment was shown to improve fatigue behavior by preventing the MnS inclusions from
fracturing during forging. In addition, the grain flow and inclusion redistribution resulting
from forging enhanced the mechanical properties of the base material
Chastel et al [12] performed an analysis of the impact of forging on fatigue of
steels. Their study involved a finite element model for analyzing the forging process.
This model considers grain flow and anisotropy. Figure 2-3 shows the grain flow of a
steel part. This grain flow can also lead to grain fragmentation and deformation of
inclusions. Ductile inclusions can stretch. Hard inclusions are realigned. This results in
anisotropic fatigue properties. Once the anisotropic mechanical properties are determined
the authors suggest using the Murakami model, which predicts fatigue limit for
specimens with defects and inclusions. The Murakami equation treats defects and

8
inclusions as cracks. The following equation is for determining the stress intensity factor
at the tip of a surface defect or inclusion:
area K t o A = A 65 . 0 (2.1)
The area parameter is defined as the defect area projected onto the plane perpendicular
to the applied stress (see Figure 2-4). The following equation relates the threshold stress
intensity factor range,
th
K A , to the defect size:

4
) 10 ( HV 226 . 0
3 / 1 3
2
5 . 0 ) )( 120 HV ( ) 10 ( 3

+

+ = A
R
area K
th
(2.2)
In this equation R is the stress ratio and HV is the Vickers hardness. When Eqns. 2.1 and
2.2 are combined, the result is:

4
) 10 ( HV 226 . 0
6 / 1
2
5 . 0
120 HV
43 . 1

+
(


+
=
R
area
w
o (2.3)
where
w
o is the fatigue limit. It should be noted that Eqn. 2.3 is for surface defects. For
internal defects the constant in Eqn. 2.3 of 1.43 is changed to 1.56. The projected area of
the inclusion perpendicular to the normal stress is used to characterize the anisotropic
effects of the deformed inclusion.

2.2 Surface Finish
It is widely recognized that surface finish has a significant effect on fatigue
behavior. Forging is usually accompanied by considerable surface roughness, surface
decarburization and scale defects. In general, fatigue life decreases as surface roughness
increases, particularly in the high cycle fatigue (HCF) region. Surface roughness has less
of an effect in the low cycle fatigue (LCF) region. Decarburization results from forging

9
process as well as heat treatment. It can be described as a reduction of carbon starting
from the surface. Decarburization results in a reduction of hardness in the decarburized
area. Surface cleaning treatments such as sand blasting decrease the effect of the as-
forged surface finish by removing scale defects, as well as some of the decarburized
layer, and to some extent inducing compressive residual stresses at the surface.

2.2.1 Decarburization
Hankins and Becker [3] tested forged specimens of different hardness under
cantilever type rotating bending on four types of forged steels, as listed in Table 2.2. The
as-forged surface finish specimens had an hourglass shape with a 0.3125 inch (7.94 mm)
test section diameter. Machined and polished specimens had an hourglass shape with a
0.25 inch (6.35 mm) diameter. Half of the specimens were tested in the as-forged
condition. It should be noted that the flash had been trimmed before testing. The other
half had the as-forged surface removed prior to testing. A metallurgical analysis of the
test specimens showed that the hardness of the decarburized surface was lower than the
hardness of the inner material (see Table 2.2). There is less of a difference between
surface and core hardness as specimen hardness decreases. The specimens with the
largest difference between surface and interior hardness had the largest difference in
endurance limit between the polished and as-forged surface conditions (see Table 2.3 and
Figure 2-5). The authors state that decarburization is the main cause of reduced fatigue
life for specimens of the as-forged surface condition.
Hankins et al [4] then performed further experiments on the effects of surface
conditions on fatigue resistance of steels. In their study high tensile strength forgings

10
were subjected to rotating bending fatigue. The fatigue specimens had the same
dimensions as in their previous study [3]. The high tensile strength as-forged specimen
had a very low endurance limit when compared to the machined and polished condition
(see Table 2.4 and Figure 2-5). Normally the endurance limit of a polished specimen is
around half of the ultimate tensile strength. However, in Table 2.4, it can be seen that the
endurance limit of the polished specimens is only around 40% of the tensile strength. As
a result, the author decided to subject the machined specimens to a better polishing
procedure, which resulted in a significant improvement in endurance limit (683 MPa
increased to 896 MPa). The authors suggest that the surface irregularities may have more
of an effect on fatigue strength than the surface decarburization for high tensile strength
forgings.
In a paper by Noll and Lipson [2], as-forged specimens of different hardness
levels were investigated. It was found that the surface hardness was much lower than the
hardness of the inner material for the forged specimens due to the decarburization, similar
to results from Hankins et al. Figure 2-6 shows plots of the Brinell hardness versus
distance form the surface for specimens of four hardness levels. Figure 2-7 shows
magnified photographs of the specimens cross-section showing the decarburized layer. It
should be noted that the pictures in Figure 2-7 correspond to the plots in Figure 2-6. The
lighter colored area shown in the pictures represents the decarburized section. It can be
seen in Figure 2-6 that the surface hardness was significantly lower than the interior
hardness. This difference in hardness is one explanation for decreased life in the HCF
region for forged specimens, since higher hardness (higher strength) is a desired property
for long fatigue life. As previously mentioned, controlled forging can result in less

11
decarburization and increased endurance limit, as seen in Figure 2-8, which shows a plot
of endurance limit versus tensile strength for as-forged specimens. There are two curves
in this figure, with the upper curve representing a controlled forging process and the
lower curve representing a less controlled (standard large scale manufacturing) forging
process [13]. The controlled forging process results in decarburization depths less than
0.13 mm, compared to depths as deep as 0.89 mm for the less controlled process [13]. In
addition, the controlled forging process results in less surface irregularities than the
standard large scale forging process.
Gildersleeve [14] investigated the relationship between decarburization and
fatigue strength of a low alloy steel (605 M36). Rotating bending tests were used to
determine the fatigue behavior. The specimens had decarburized layers up to 1 mm in
depth. The results showed that the fatigue limit was mostly independent of the depth of
decarburization (see Table 2.5). The author also examined surface carbon concentration
and found the fatigue limit to be linearly dependent upon the carbon concentration at the
surface.
Adamaszek and Broz [15] investigated the effects of decarburization on hardness
changes in carbon steels caused by high temperature surface oxidation. The
decarburization they studied resulted from annealing. They state that the decarburization
causes the grains near the surface to grow. In addition to grain growth, there is formation
of surface scales, which are solid, firm and porous. The authors explain that during the
decarburization process oxygen penetrates the surface through cavities, pores and cracks.
This oxygen reacts with the different chemicals (elements) in the metal causing the
decarburization. The decarburization is worse for metals with higher Fe concentration.

12
The authors also state that there are hardness changes due to the decarburized layers
resulting in lower fatigue resistance.

2.2.2 Roughness measurements and parameters
Before determining how surface roughness affects fatigue life, it is necessary to
measure and define surface roughness. The most common method of measuring surface
roughness is the mechanical profiler. It works by dragging a stylus (probe) across the
surface [16, 17]. As the stylus is drawn slowly across the surface, it moves up and down
with the contours of the surface. This motion is then recorded. However, this instrument
is limited by the radius of the stylus tip. If the radius is too large it may not be able to
penetrate the finer cracks or scratches, resulting in an incorrect surface roughness
measurement. It is also a possibility that the surface could be damaged by the stylus. A
non-destructive method of measuring surface roughness is the laser speckle contrast
method [16]. In this method a helium-neon light is pointed at the surface at various
angles. As this light hits the surface it is reflected creating a speckle. It was found [16]
that there is a linear relationship between the surface roughness and the speckle contrast
of an illuminated surface up to a certain roughness (0.1 m R
a
).
Once the surface topography is recorded it is necessary to define it. Gadelmawla
et al [18] describe 59 different parameters for describing surface roughness. The authors
state that the arithmetic average height parameter (R
a
), also known as the center line
average (CLA), is the most widely used parameter. This parameter is the average
deviation from the mean line over a sampling length (see Figure 2-9(a)). This is shown
mathematically below:

13
}
=
l
a
dx x y
l
R
0
) (
1
(2.4)
While R
a
is easy to define, it does not describe wavelength and is not sensitive to small
changes in roughness profile. It should be noted that when reporting the value of R
a
it is
necessary to also report the cutoff length, which is the length that the roughness is
averaged over. R
q
or RMS is the root mean square roughness, which is the standard
deviation of the distribution of surface heights, and is another common parameter. R
q
is
more sensitive to a large deviation from the mean line than R
a
. This parameter is
expressed mathematically as:

}
=
l
q
dx x y
l
R
0
2
)} ( {
1
(2.5)
The parameter R
z
, known as the ten-point height, is defined as the average
summation of the five highest peaks and the five lowest valleys (see Figure 2-9(b)). It
should be noted that R
a
is not as sensitive to occasional peaks and valleys as R
z
. R
z
is
defined mathematically by:
|
.
|

\
|
+ =

= =
n
i
n
i
i i DIN z
v p
n
R
1 1
) (
1
(2.6)
The maximum height of profile parameter, R
t
, is most sensitive to large peaks and
valleys. It is defined as the distance measured between the highest peak, R
p
, and the
lowest valley, R
v
, (see Figure 2-9(c)). It is defined mathematically by:

v p t
R R R + = (2.7)
In Figure 2-9(c), R
v
is equal to R
v4
and R
p
is equal to R
p3
. Therefore, R
t
is equal to R
p3
+
R
v4
. In Figure 2-9(c), R
y
, not to be confused with R
t
, is the largest peak to valley height
(i.e. R
p3
+ R
v3
).

14
Mean spacing between peaks, S
m
, is another parameter used to characterize
surface roughness. It is defined as the mean spacing between profile peaks at the mean
line (see Figure 2-9(d)) and is defined mathematically as:

=
=
n
i
i m
S
N
S
1
1
(2.8)
There are many more parameters used to quantify surface roughness. Each parameter is
designed to be more sensitive to different variations in roughness such as height and
depth of roughness, frequency of roughness, and distance between peaks.
Novovic et al [19] performed a literature review on the effect of machined surface
topography and integrity on fatigue life and examined different roughness parameters.
They examined surface roughness parameters and concluded that R
a
is the most
commonly used parameter in describing fatigue behavior. However, they found that there
is typically a 20% scatter in fatigue results for specimens of the same R
a
value. The
authors suggest that R
t
and R
z
are better to use in determining fatigue performance than
R
a
, because these parameters represent the worst defects in the surface.

2.2.3 As-forged surface condition
As stated earlier, most of the available data for the effect of as-forged surface
condition on fatigue are old and very conservative. The paper entitled Allowable
Working Stresses by Noll and Lipson [2] is one of the main sources of data used to
develop the endurance limit modification factors for the as-forged surface condition,
which are still used in fatigue design and analysis. The authors investigated the
relationship between endurance limit (defined at 10
7
cycles to failure) and surface
condition (ground, machined, hot-rolled, and forged). The forged surface condition is

15
described as having large surface irregularities, including oxide, scale defects, as well as
total surface decarburization and is regarded as the worse of the above mentioned surface
conditions. They grouped fatigue data on steels with hardness ranging from 160 HB to
555 HB from previous studies by other authors into the above categories. Some of the
data for the as-forged surface condition are from data published by Hankins et al [3, 4].
However, most of the data for the as-forged surface condition were obtained from
fabricated parts tested at the Chrysler Laboratories. The actual test data and experimental
details performed at Chrysler Laboratories, from which the data were derived, are not
included in [2]. From this data, Noll and Lipson developed several figures of allowable
stress versus mean stress for each type of specimen (see Figure 2-10 for an example).
Data from the 1945 version of the SAE Handbook were used by Noll and Lipson to
define the relationship between tensile strength and hardness (see Figure 2-11), which is a
mostly linear relationship. Noll and Lipson used the most conservative values from
Figure 2-11 in their analysis. In addition, they developed a figure plotting endurance limit
versus tensile strength (see Figure 2-12) for the four surface conditions investigated. The
data used to develop Figure 2-12 are listed in Table 2.6. It can be seen in Figure 2-12 that
the as-forged surface condition results in the lowest endurance limit for a given tensile
strength of the four surface conditions described in [2]. It should be noted that controlled
forging conditions can produce surface finish quality similar to hot-rolled surfaces [13].
In a discussion of the paper Allowable Working Stresses, Lessells compares
endurance limit data on as-forged and ground and polished surface conditions from Noll
and Lipson to data from Hankins et al [3, 4]. Lessells found the data from Noll and
Lipson to be much more conservative than the data from Hankins et al. Noll and Lipson

16
in their reply state that the discrepancies are due to modification of data for size effect,
reducing Hankins endurance limit data by 15%. Noll and Lipson explain that there is
more of a difference between data for forged surface than for the ground and polished
surface due to the broad definition of the forged surface condition.

2.2.4 Effect of surface roughness on fatigue behavior
Fluck [20] studied the influence of surface roughness on fatigue life and scatter of
test results of two steels. The steels used for the cantilever rotating bending tests were a
quenched and tempered SAE 3130 steel (30 HRC) and an annealed SAE 1035 steel (69
HRB). The specimens, which were machined from 12.7 mm rolled bars, were grouped
into six categories of surface roughness, lathe-formed, partly hand-polished, hand-
polished, ground, ground and polished, and superfinished. The surface roughness
measurements were made with a Brush surface analyzer. This instrument indicated the
root mean square of the surface roughness. Figure 2-13 shows a plot of the RMS (R
q
)
surface roughness versus median fatigue life. The results show that the ground and
polished specimens, which represent the smoothest surface, always experienced the
longest life and the lathe-formed specimens, which represent the roughest surface,
experienced the shortest life.
Figure 2-14 shows the scatter of these results. It can be seen in Figure 2-14 that
there is generally more scatter for a given polishing condition at longer life for both of the
materials used. The author concluded that fatigue life can be significantly increased by
reducing the size of circumferential scratches. Specimens polished to roughness below
six microinches experienced a large increase in fatigue life. In a discussion of this paper,

17
Lessells points out that the author did not consider surface residual stresses and surface
hardening through cold work and martensite formation that could have been introduced
by the polishing procedures. As a result, according to Lessells, surface roughness is not
the only cause for the differences in fatigue life data reported by Fluck.
Sibel and Gaier [21] investigated the influence of surface roughness on fatigue
strength of steels and non-ferrous alloys [21]. Axial and reverse bending fatigue tests
were performed on several types of steels (medium carbon, Cr-Mo, spring steel, and
stainless) as well as brass and some aluminum alloys. Roughness was applied to the
specimen surface by polishing, grinding and turning. The specimens had an hour glass
shape with 8.4 mm and 7.5 mm minimum diameters for the axial and bending specimens,
respectively. The roughness parameter used was R
v
, the maximum depth of groove, with
values ranging from 1 to 50 microns. There was a similar decrease in fatigue strength, for
both axial and reverse bending fatigue tests, as surface roughness increased.

2.2.5 Fatigue limit evaluation for surface finish effect
Murakami and Endo [22] performed an extensive literature review on the effects
of defects, inclusions and inhomogenities on fatigue strength and the existing models.
The authors classify the different approaches into three categories: empirical models,
models based on fatigue notch factor approach, and fracture mechanics models. Many of
them can be used to determine behavior of specimens with scratches, cracks or notches.
As presented earlier, Murakami et al developed an equation (Eqn. 2.3) for
prediction of the fatigue limit for specimens with small defects and inclusions [22].
Although this equation was developed for specimens with small defects and inclusions,

18
they have shown it to be applicable to surface roughness conditions by performing
rotating bending fatigue testing on a medium carbon JIS S45C steel [23, 24]. The
specimens were machined after heat treatment or annealing (see Figure 2-15 for specimen
configuration and dimensions). After machining, artificial roughness in the form of
notches of various depth and pitch was applied to the specimens by use of a lathe. The
depth of the artificial roughness was considered random because of build up on the
cutting tool, but the pitch between notches was considered constant (see Figure 2-16 for
magnified views of the different types of artificial surface roughness applied). There were
also some electro-polished specimens. Some of the electro-polished specimens had a
single notch applied to them after polishing. Roughness was measured by use of a
mechanical profiler. Figure 2-17 shows a plot of the stress-life data. All fatigue failures
occurred at the root of an artificially induced notch. The fatigue strength decreased as the
depth of the notch increased. It was shown that the specimens with a single notch
experienced fatigue strength that was about 30% lower than that of specimens with
multiple notches. This is because interference between notches reduces the fatigue notch
effect. As a result the authors determined that the pitch between notches must be
considered, which is not considered in the area parameter.
In order for Eqn. 2.3 to be used to evaluate surface roughness, Murakami et al
developed an equivalent defect size for surface roughness
R
area to replace area ,
which accounted for both depth and pitch. They assumed that periodic roughness notches
are equivalent to periodic cracks. It should be noted that this problem was evaluated as a
crack problem and not a notch problem. They derived the following equations:

19

3 2
) 2 / ( 47 . 9 ) 2 / ( 51 . 3 ) 2 / ( 97 . 2
2
b a b a b a
b
area
R
~ For 195 . 0 2 / < b a (2.9)
38 . 0
2
~
b
area
R
For 195 . 0 2 / 3 > > b a (2.10)
where a is the crack depth and 2b is the distance between the two cracks. Figure 2-18
shows a plot of
R
area /2b versus a/2b. This figure includes superimposed data points
calculated from experimental data. When considering value of depth, a, the authors chose
to use the maximum height parameter R
y
. The following equation was used to calculate
the fatigue limit for the electro-polished specimens (without notch) and the artificially
roughened specimens, respectively:
HV 6 . 1 =
w
o ) 400 HV ( s For polished surface condition (2.11)

6
1
) (
] 2 / ) 1 )[( 120 HV ( 43 . 1
R
w
area
R
o
o
+
= For rough surface condition (2.12)
Table 2.7 displays the
R
area values for each type of artificial surface roughness. Table
2.8 shows the experimental and calculated values of fatigue limit along with the hardness
values and the
R
area parameter. Materials 100A and 150QT listed in Tables 2.7 and
2.8 have similar crack depth, but different pitch (100 m, 150 m and 200 m). The
numbers in the specimen ID listed in Tables 2.7 and 2.8 refer to the pitch. It can be seen
that the decreased pitch distance results in a lower value of
R
area . This lower value of
R
area would result in a larger value of
w
o , which is in agreement with experimental
data. Figure 2-19 shows a plot of the experimental life versus the predicted life using
Eqns. 2.11 and 2.12, showing an error less than 10% in predicting fatigue limit for most
specimens.

20
Shareef and Hasselbusch [5] investigated endurance limit modifying factors for
hardened machined surfaces. They state that existing surface finish modification factors,
derived from Allowable Working Stresses, are limited to hardness up to 33 HRC, and
when this factor is extrapolated to higher hardnesses it produces overly conservative
values. The authors performed fatigue tests on specimens made from SAE 4140 steel
with a machined surface and hardness ranging between 50 and 55 HRC. The average
hardness of the specimens tested was 53.4 HRC. The roughness was defined by the R
a

parameter. The specimens had an average roughness of 3.51 m. The specimens also had
surface compressive residual stresses with an average value of -331 MPa. Figure 2-20
shows the specimen geometry as well as the test set up used. The shafts were subjected to
three-point bending while strain was monitored using gages. They used three different
stress equivalency methods to calculate the surface finish correction factor, k
a
:

polished
e
a
a
machined
e
a
S
k
S
|
|
.
|

\
|
=
|
|
.
|

\
| o o
(2.13)
( ) ( )
c
f f
b
f
f a
N N
E
k
2 2
2
0
c
o o
c
' +
|
.
|

\
|

'
=
|
.
|

\
| A
(2.14)

( )
( ) ( )
c b
f f f a
b
f
f a
N k N
E
k
+
' ' +
'
= |
.
|

\
| A
2 2
2
2
2
max
c o
o
c
o (2.15)
They solved for k
a
and found that the SWT relationship (Eqn. 2.15) resulted in the best
predictions. Table 2.9 shows the correction factors calculated using Eqns. 2.13-2.15 for
10
4
and 10
6
cycles to failure. It should be noted that in Eqns. 2.14 and 2.15, k
a
was
applied to only the elastic portion of the curve. Figure 2-21 shows that Eqns. 2.14 and
2.15 result in a decreased effect of surface roughness with a decrease in fatigue life, but
Eqn. 2.13 shows no difference between LCF and HCF. Existing data, derived from

21
Allowable Working Stresses, gives an endurance limit modification factor of k
a
= 0.44
for machined specimens at 53.4 HRC. Shareef and Hasselbusch state that the use of their
most conservative modification factor k
a
= 0.75, derived from Eqn. 2.13, would result in
cost and weight savings up to 50%. This is an example of how conservative the data
published by Noll and Lipson are.
Arola and Williams [25] investigated the effects of surface texture on a high
strength low alloy steel. In their investigation, surface roughness parameters were used to
calculate an effective stress concentration factor, which was then used to determine an
effective fatigue notch factor. The effective stress concentration factor based on
roughness parameters was defined as:
|
|
.
|

\
|
|
|
.
|

\
|
+ =
z
y
a
t
R
R
R
n K

1 (2.16)
where R
a
is the average roughness, R
y
is the peak to valley height, R
z
is the 10-point
roughness, and is the average radius determined from the dominant profile valleys (see
Figure 2-22 for definition of ). The value of n in Eqn. 2.16 is equal to two for tension
loads and equal to one for shear loads. Equation 2.16 can then be substituted into Eqn.
2.17 below to calculate an effective fatigue notch factor, as:
) 1 ( 1 + =
t f
K q K (2.17)
In Eqn. 2.17, q is defined as the notch sensitivity and is given by:

|
.
|

\
|
+
=

1
1
q (2.18)
where for steels is define by:

22

8 . 1
2070
025 . 0
|
|
.
|

\
|
=
u
MPa
o
(2.19)
In Eqn. 2.19,
u
o is the ultimate tensile strength in MPa and is in units of mm. Figure
2-23 shows a plot of fatigue stress concentration factor versus average surface roughness.
This plot superimposes calculated values along with experimental values. It can be seen
that the proposed equation for stress concentration factor is in good agreement with the
experimental data.
Stephens et al [26] state that the fatigue notch factor can be applied to the
modified Goodman equation as follows:
1 = +
|
.
|

\
|
u
m
f
f
a
S
S
K
S
S
(2.20)
where S
a
is the net section alternating stress, S
f
is the fully-reversed fatigue strength for a
smooth unnotched specimen, S
m
is the mean stress, and S
u
is the ultimate tensile strength.
Therefore, it is then possible to apply the effective fatigue notch factor calculated from
Eqn. 2.17 to the modified Goodman equation to account for mean or residual stresses,
stress concentration, and surface finish effects.
There are many textbooks and handbooks that describe how surface finish affects
fatigue life. The methods of evaluating the as-forged surface condition presented in them
give conservative values. Mechanical design and fatigue textbooks generally provide an
equation for determining the endurance limit for the surface finish condition (for example
in [27]), as well as other factors, by use of correction factors, given by:

e f e d c b a e
S k k k k k k S ' = (2.21)

23
In this equation k
a
is the surface modification factor, k
b
is the size modification factor, k
c

is the load modification factor, k
d
is the temperature modification factor, k
e
is the
reliability factor, k
f
is the miscellaneous-effects modification factor, and
e
S' is the
endurance limit from rotating bending fatigue tests performed on smooth polished
specimens. The surface modification factor is defined as:

b
ut a
S a k = (2.22)
In this equation
ut
S is the ultimate strength and a and b are constants found in Table 2.10.
Table 2.10 provides values of a and b for the ground, machined, hot-rolled and as-forged
surface conditions. These values are derived from the paper Allowable Working Stresses
[2] discussed earlier. Figure 2-24 shows a plot of surface finish factor versus tensile
strength presented in many mechanical design and fatigue books [such as 26-34] which is
derived from the data presented in Figure 2-12. The authors of Allowable Working
Stresses determined endurance limit at 10
7
cycles when developing Figure 2-12 [2].
However, it is suggested by Noll and Lipson [13] that steels reach the endurance limit
value around 10
6
cycles, as seen in Figure 2-25, which shows a plot of stress versus
cycles to failure for as-forged surface finish specimens of two hardness levels [13]. It is
mentioned that the surface finish factor has an increasing effect with an increase in tensile
strength or hardness, which can be seen in Figure 2-24. In addition, the surface finish
effect is generally thought to be negligible for specimens subjected to less than a
thousand cycles due to a high degree of plastic deformation. Figure 2-26 shows an
example of S-N behavior for specimens having machined and polished surface and as-
forged surface conditions. It can be seen that the curves intersect at 10
3
cycles.

24
The load factor, k
c
in Eqn. 2.21, which accounts for the type of loading the
specimen is subjected to is given by [27]:

=
59 . 0 577 . 0
85 . 0
1
c
k
torsion
loading axial
bending rotating
(2.23)
The correction factor for rotating bending is one because
e
S' is the endurance limit under
rotating bending loading, therefore no correction is needed.
Size effect modification factor, k
b
, corrects for diameters larger than about 7 mm,
since a standard diameter of a test specimen is about 7 mm [27]. For specimens between
12 mm and 25 mm diameter, there is around a 15% reduction in fatigue limit under
rotating bending loading. For specimens larger than 50 mm in diameter, the fatigue
strength can be 25% lower than the standard specimens fatigue strength under rotating
bending loading. The size factor is given mathematically by:

( )

s <
s s
=

mm 254 51 51 . 1
mm 51 79 . 2 762 . 0 /
157 . 0
107 . 0
d d
d d
k
b
(2.24)
For nonrotating bending the size effect is less significant. This is because there is a lower
volume of material subjected to maximum stresses under nonrotating bending. The lower
volume of material is accounted for by using an effective diameter in Eqn. 2.24, given by:
d d
e
37 . 0 = (2.25)
The difference between rotating and nonrotating bending fatigue strength is regarded as
typically less that 5% for commercial materials [35].




25
2.3 Sand Blasting and Shot Cleaning
Sand blasting is a process in which sand is shot, by air, at the surface of a
component for the purpose of cleaning the surface. It is an effective method of removing
the scale left from forging and heat treating operations. There are other methods of
treating the surface such as shot peening, nitriding and other chemical processes, but
these surface enhancement processes are outside the scope of this study. The sand
blasting process induces compressive residual stresses at the surface as well as increased
surface roughness. Sand blasting typically produces compressive layers between 25 and
75 m deep compared to compressive layers between 100 and 300 m resulting from the
shot-peening process [19]. The depths of the compressive layers for sand blasting
mentioned are, however, for a well-controlled sand blasting process in which the
compressive stresses are intentional. This is not to be confused with a sand blasting
process whose intentions are to clean the surface, which also typically leaves
unintentional beneficial compressive residual stresses. Typically the sand blasting
treatment for steels increases fatigue life in the HCF region due to the beneficial
compressive residual stresses, but may decreases fatigue life in the LCF region due to
reduced surface ductility.
Hanley and Dolan [36] conducted a literature review on surface finish and its
effect on fatigue behavior. They refer to a study by Wiegand [37], which concluded that
shot blasting increased the fatigue strength (at 10
7
cycles) of a smooth surface by 10 MPa
as well as a roughened surface by 90 MPa. Wiegand also found that sand blasting
decreased fatigue strength (at 10
7
cycles) of a polished surface by about 48 MPa. In a
study by Manteuffel [38] sand blasting was shown to improve fatigue strength (at 10
7


26
cycles) of steel springs. Hanley and Dolan explained that the differences in fatigue data
from [37] and [38] were due to the initial surface conditions, polished and unpolished.
Zimmerli [39] subjected shot blasted springs to fatigue testing and found that fatigue
strength (at 10
7
cycles) was improved by the shot blasting process. He also found that a
smaller sized shot (0.4 mm diameter) improved long life fatigue behavior for a given
maximum stress, compared to a larger shot (1.2 mm diameter). In addition, he found that
light sand blasting following the shot blasting further increased fatigue life for a given
maximum stress. However, neither the shot blasting (0.4 mm shot) nor shot blasting (0.4
mm shot)/sand blasting increased the fatigue strength at 10
7
cycles compared to the larger
shot, it only increase fatigue life for a given stress level.
Farrahi et al [1] studied the influence of residual stress on the fatigue life of
forged and shot blasted components. In their study, hot forged round EN15R steel bars
were subjected to uniaxial tension-compression fatigue tests. Specimens were forged by
use of induction heat and cooled in room temperature air. The forged specimens had an
hourglass shape with a minimum diameter of 8.2 mm. Some of the forged specimens
were heat treated and/or sand blasted. The flash line had been trimmed prior to heat
treatment and shot blasting. The sand blasting was done to clean the scale off the
specimen as well as to induce compressive residual stresses. Results showed that sand
blasting was beneficial in the HCF region due to compressive residual stresses (see Table
2.11). However, sand blasting was found to be detrimental in the LCF region due to
relaxation of the compressive residual stress resulting from plastic deformation in the
LCF region and reduced ductility due to plastically deformed surface. The sand blasting
treatment increased scatter of test results in both LCF and HCF regions.

27
2.4 Heat Treatment and Hardness Effects
2.4.1 Heat treatment
Heat treatment of steels, which is done to control material properties, to a large
extent is dependent on carbon content adding strength to the material [40-43]. Iron can
take two crystallographic orientations, BCC and FCC, which can be seen in Figure 2-27.
The iron changes from one atomic structure to another when it is heated to a certain
temperature. The solubility of carbon is higher for the FCC orientation. By controlling
the rates of heating and cooling of the material, properties of the material can be
controlled. Slowly cooling the material allows the carbon to leave the metal. However
rapid cooling traps the carbon in the material making it harder. It should be noted that this
is a reversible process.
Quenching and tempering is a common heat treatment procedure. Table 2.12 lists
a variety of other heat treatment procedures for steels along with their purpose. The idea
behind quenching and tempering is to quench the material when it has reached the
austenite phase (see Figure 2-28, a phase diagram plotting the temperature versus the
carbon concentration). It can be quenched in different liquids such as water or oil. Once
quenched, the material rapidly cools and martensite forms. Figure 2-29 shows a plot of
hardness of martensite versus carbon content, indicating the hardness increases with
increasing carbon content. Following quenching, it is necessary to temper the material
because the quenched material is brittle and contains internal stresses. The internal
stresses are caused by expansion and contraction of the material during the heating and
cooling process.

28
Tempering is done by reheating the material to various temperatures to obtain
desired properties as well as relieve internal stresses [40-43]. Figure 2-30 shows a plot of
tensile and yield strength versus the tempering temperature for an oil-quenched 4340
steel having a martensitic microstructure [43]. This figure shows that tempering at lower
temperatures allows the material to maintain strength or hardness while relieving internal
stresses. As tempering temperatures increase the ductility of the material increases but the
strength or hardness decreases. Some side effects of quenching and tempering are
cracking and decarburization. Cracking can be caused by quenching from too high a
temperature or by non-uniform heating and cooling. Decarburization is caused by an
oxidizing atmosphere similar to decarburization resulting from the forging process,
already discussed.
The Steel Heat Treatment Handbook [41] describes how undesirable distortion
can occur during the heat treatment process. Material properties will affect thermal
expansion of the material, which if not accounted for can cause residual stresses to form
during quenching. The homogeneity of the material should be considered. Large
variations of composition across the cross section of the material can cause unwanted
residual stresses during heat treatment. The residual stresses induced by processes such as
forging or hot rolling can be relieved during the heat up portion of the heat treatment
process. However, when these stresses are released, it can cause the part to deform. There
are also shape changes due to phase changes during heat treatment. There is about a 4%
reduction in volume when iron transforms from the pearlite phase to the austenite phase.
As a heat treated part cools, residual stresses can form, mostly due to phase changes.
However, residual stresses can be minimized by slowing down the cooling rate.

29
2.4.2 Hardness effects on fatigue behavior
Hankins et al investigated the fatigue resistance of unmachined forged steels [3,
4]. Results of the monotonic tension testing show that as hardness increases so does the
tensile strength. Their fatigue test results show that the harder the material the greater the
difference in endurance limit between as-forged and polished specimens. This can be
seen in Figure 2-5, which shows a plot of endurance limit (at 10
7
cycles) versus tensile
strength for as-forged and polished materials tested in [3, 4]. The fatigue limit for the as-
forged surface is up to 83% lower than the polished surface for the hardest material, but
only 15% lower for the softest material tested (see Tables 2.3 and 2.4). It should be noted
that the softest material was not quenched and tempered, which was not the case for the
other materials. As mentioned in the section on decarburization (section 2.2.1), Hankins
et al attribute decarburization to the decrease in endurance limit for the as-forged
specimens of hardness below 308 HB. They also state that the surface condition is of
more importance to materials of higher hardness, as in the case of the two high strength
steels tested (listed in Table 2.4).
In Allowable Working Stresses, by Noll and Lipson [2], they grouped data from
other investigators into different types of surface condition with different hardnesses
ranging from 160 HB to 555 HB. It was shown that the surface condition had an
increasing effect with increased hardness, which can be seen in Figure 2-12 in a plot of
endurance limit versus tensile strength. This plot shows that there is little to no increase
in endurance limit for a given tensile strength for the as-forged surface condition. Fluck
[20] found similar results when investigating surface roughness effects on two steels,
SAE 3130 (30 HRC) steel and an annealed SAE 1035 (69 HRB) steel. It was also shown

30
by Noll and Lipson that a linear relationship exists between tensile strength and Brinell
hardness (see Figure 2-11).
Landgraf [44] investigated cyclic deformation and fatigue behavior of hardened
steels. SAE 1045 and SAE 4142 steels were used in their study. The specimens were
machined and then heat treated to five different hardness levels ranging from 380 HB to
670 HB. Strain-controlled completely-reversed axial fatigue tests were performed. Step
tests for cyclic deformation and monotonic tension tests were also performed. The results
of monotonic testing showed a linear increase in true facture strength with an increase in
hardness up to 600 HB. After 600 HB, the true fracture strength began to decrease. Figure
2-31 shows superimposed plots of monotonic and cyclic stress-strain curves for the
different hardnesses for the SAE 1045 and SAE 4142 steels. It can be seen that most
specimens cyclically softened and the degree of cyclic softening decreases as hardness
increases. However, at the two highest hardness levels for the SAE 4142 steel (670 HB
and 560 HB) the material cyclically hardened. Figure 2-32 shows plots of stress
amplitude versus plastic strain amplitude for the different types of specimens used. This
figure shows very little effect on the cyclic strain hardening exponent with changes in
hardness; however there is an increase in cyclic strength coefficient with an increase in
hardness. Figure 2-33 shows plots of total strain amplitude versus reversals to failure for
SAE 1045 steel at 700 HB, 600 HB, and 450 HB, as well as SAE 4142 steel at 670 HB,
560 HB, and 450 HB. Figure 2-34 shows plots of total strain amplitude versus hardness at
different cycles to failure for SAE 1045 steel and SAE 4142 steel. It can be seen in these
figures that the softer materials show better fatigue resistance in the LCF region due to

31
the increased ductility. In contrast, the harder materials show better fatigue resistance in
the HCF region due to increased strength.

2.5 Crack Growth Behavior
2.5.1 Fracture mechanics
Stress-life and strain-life analysis do not distinguish between crack initiation and
crack growth. Typically, for smooth unnotched fatigue specimens, most of the life is
taken up by crack initiation. However, for parts having a rough surface, micro-cracks
may exist prior to loading depending on the severity of the roughness. The majority of
life for a part containing surface cracks will consist of crack growth. Fracture mechanics
characterizes fatigue crack growth behavior by use of a stress intensity factor or the J-
integral. There are two types of fracture mechanics analysis, linear elastic fracture
mechanics (LEFM) and elastic-plastic fracture mechanics (EPFM). The type of fracture
mechanics analysis used is dependent upon crack-tip behavior. Although the nominal
stress applied to a part may be elastic, the stress concentration resulting from the crack
can cause significant plastic deformation at the crack tip. LEFM may be used when the
plastic deformation at the crack tip is small, and this is the case in most fatigue problems.
EPFM is used when there is significant plastic deformation at the crack tip [26].
A LEFM approach to modeling crack growth behavior under mode I (i.e. tensile
mode or opening crack mode) which is most common in applications is applied by
plotting crack growth rate, da/dN, versus stress intensity factor range, K, in a log/log
plot, where:
o t a S K A = A (2.26)

32
In this equation, S is the stress range (zero stress to maximum stress if the minimum
stress is compressive), a is the crack length, and is a geometry factor dependent upon
crack geometry and loading mode. The change in crack length divided by the change in
cycles (da/dN) is the rate at which the crack grows. Figure 2-35, which is for the linear
Paris equation regime with crack growth rates higher than 10
-8
m/cycle, shows crack
growth rate versus stress intensity factor range for a wide range of martensitic steels
having tensile strengths ranging from 730 MPa to 1290 MPa. The Paris equation is given
by:

n
K A
dN
da
) (A =
(2.27)
where A is the intercept at K = 1, and n is the slope. It can be seen in this figure that
there is relatively little scatter for a type of steel having the same microstructure with data
falling in between two narrow scatter bands. This suggests that the applied stress intensity
factor range is the governing factor controlling crack growth behavior in steel and not
material properties (i.e. yield strength, tensile strength and ductility) [45]. Use of the
upper scatter band equation would result in conservative predictions of crack growth
rates.

2.5.2 Fatigue crack growth
Arieli and Mukherjee [46] performed tension-tension cyclic tests on notched
specimens in order to determine the effects of decarburization on fatigue life (see Figure
2-36 for specimen dimensions). The material used was a 4340 low alloy steel with an
ultimate tensile strength of 965 MPa. They stated that residual tensile stresses may be
present in the decarburized layer due to volumetric contraction of the low carbon area by

33
the higher carbon content of the inner material. Test results showed that fatigue life
improved in the LCF region for decarburized specimens compared to specimens without
decarburization. As the decarburization depth increased, so did the fatigue life for a given
stress amplitude, as shown in Table 2.13, which lists fatigue life for various
decarburization depths, where a
d
indicated depth of decarburization. However,
decarburization eliminated the possibility of run-out. This behavior is believed to be due
to crack growth behavior of the decarburized layer [46]. The authors also noted that
degree of fatigue life improvement was dependent upon the maximum applied stress and
not the stress amplitude. The fatigue test results listed in Table 2.13 include the total
fatigue life (N
t
), the fatigue life for crack to grow through the decarburized layer (N
d
), and
fatigue life for crack to grow from the end of the decarburized layer to the critical length
(N
b
). Therefore, the total fatigue life was defined by:
b d t
N N N + = (2.28)

( ) ( )
( )( )
3
10
5 . 0 5 . 0
0
10 8 . 1 t o A

=


d
d
a a
N (2.29)

( ) ( )
( )( )
25 . 2
8
125 . 0 125 . 0
10 0825 . 0 t o A

=


cr d
b
a a
N (2.30)
Values of a
0
and a
d
are also listed in Table 2.13. It should be noted that Eqns. 2.29 and
2.30 were obtained by integrating the Paris equation.
Maiya et al [47] investigated the effect of surface roughness on low cycle fatigue
behavior of type 304 stainless steel. In their investigation axial fatigue tests were
performed on hourglass shaped specimens in a 593 C

atmosphere. The specimens were


heat treated and then artificial surface roughness was applied to the specimens by use of a

34
lathe and silicon carbide paper of different grits. The surface roughness was defined by
an RMS value and an R
v
value. The results of roughness measurements are summarized in
Table 2.14, where it can be seen that the maximum depth of the surface groves (R
v
) is
significantly larger than the RMS value. Figure 2-37 shows a magnified image of the
surface roughness for different degrees of roughness. The authors also suggest that
surface roughness has an effect on the early stages of crack growth rate. Figure 2-38
shows the initial crack growth rate decreases as surface roughness increases. However,
the tests data showed that fatigue cracks take less and less time to initiate as surface
roughness increases (see Figure 2-39).
Deng et al [48] evaluated the effect of surface roughness on crack initiation life.
They mention that fatigue life is separated into two parts; crack initiation and crack
growth. They state that the effect of surface roughness should be limited to the crack
initiation period of fatigue life. In order to validate this, rectangular notched specimens
(see Figure 2-40 for notch geometry) were tested in a three point bending machine at a
frequency of 30 Hz with a loading ratio (P
min
/P
max
) of R = 0.05, and with a true maximum
bending stress of 800 MPa at the notch. The notch stress was calculated using finite
element analysis. The specimen hardness ranged from 180-230 HB. Three different
degrees of surface roughness, 0.03 m, 0.3 m and 1.27 m, were applied to the notch of
the specimens. It should be noted that the roughness parameter used was R
a
. Crack
initiation was determined by use of ion-sputtered film. When a crack forms in the film,
the resistance of the film was increased. Figure 2-41 shows a plot of the number of cycles
versus the average roughness for the total fatigue life, with superimposed curves for crack
initiation life and crack growth life. It can be seen in this figure that crack initiation takes

35
up most of the total fatigue life. This figure also shows a decrease in life with an increase
in surface roughness; however the differences observed in the crack initiation period are
more significant than in the crack growth period. As a result, the authors suggest that the
effect of surface roughness should be limited to the crack initiation stage of fatigue life.

2.6 Summary
Surface finish can have a significant effect on the fatigue behavior of steels. The
as-forged surface may have considerable surface roughness, surface decarburization and
scale defects. A common surface treatment for as-forged parts is shot cleaning or sand
blasting, which is done to clean the surface. This treatment can eliminate some of the
detrimental surface effects such as scale and decarburization by removing some of the
surface layer and by inducing beneficial compressive residual stresses at the surface. Shot
cleaning has been found to improve fatigue strength in the HFC region of specimens
having rough or as-forged surfaces.
Heat treating is commonly done to forged parts. The process of heat treatment can
increase strength (hardness) and lower ductility. There is a near linear relationship
between hardness and tensile strength. The endurance strength, or fatigue strength (at 10
7

cycles), increases as the hardness increases for a smooth polished surface. However, there
is less change in fatigue strength (at 10
7
cycles) as hardness increases for parts having an
as-forged surface condition. In addition, there is an increasing detrimental effect in the
LCF region resulting from heat treatment as hardness increases due to reduced ductility.
Decarburization is a gradual reduction of carbon content starting from the surface
and can result from heat treatment as well as forging. Decarburization results in a

36
reduction of hardness, or strength, in the decarburized area. There is less of a difference
between surface and core hardness as core hardness decreases for heat treated forged
specimens. Decarburization can cause a significant reduction in fatigue strength in the
HCF region. The reduction in fatigue strength increases as hardness increases. However,
as hardness increases beyond 310 HB, surface irregularities may have more of an effect
on the fatigue behavior than decarburization. There is less of an effect in the LCF region
due to large amounts of plastic deformation. In addition, there is an increasing effect on
fatigue behavior with an increase in depth of decarburization. It has been found that
decarburization can actually increase fatigue life in the LCF region due to a reduced
crack growth rate.
There are many parameters available to define surface roughness. The most
commonly used parameter is R
a
, the arithmetic average height. However, R
a
is not as
sensitive to variations in surface roughness as other parameters such as the standard
deviation of the distribution of surface heights (R
q
), the ten-point height (R
z
), or the
maximum height of profile (R
t
). It is suggested that R
z
and R
t
better characterize fatigue
behavior as they represent the worst defects in the surface. It is well established that
fatigue life in the HCF region decreases as surface roughness increases. It should be
noted that surface finish has more of an effect on a harder material.
Fatigue life is often separated into two parts; crack initiation and crack growth.
Surface roughness has been shown to have more of an effect on the crack initiation life
and it is suggested that the effect of surface roughness should be limited to the crack
initiation period. It has been found that the crack initiation life decreases as surface
roughness increases. As crack initiation life decreases, the total fatigue life also generally

37
decreases in steels. This is because steels of the same microstructure exhibit the same
fatigue crack growth behavior (or crack growth rates).
Several models have been used to evaluate surface finish effect such as the
Murakami and the Arola-Ramulu models. The Murakami model treats surface roughness
as periodic surface cracks but the Arola-Ramulu model analyzes surface roughness as a
notch problem. These models consider surface roughness and do not consider the effects
of decarburization. There are also surface finish modification factors which are
commonly applied to the endurance strength of the polished condition. However, existing
modification factors for the as-forged surface condition, which are based on data dated
back to 1930s and 1940s, have been found to be too conservative.

























38
Table 2.1: Commonly used temperatures when heating selected ferrous alloys for hot
forging [6].




Table 2.2: Hardness variation between surface and subsurface for four types of
forged steels used in fatigue tests in [3].








At Surface Interior Difference
A 150 154 4
B 146 158 12
A 206 221 15
B 171 192 21
A 233 284 51
B 215 247 32
A 201 333 42
B 242 289 47
Diamond Pyramid Hardness
3% nickel steel
nickel-chromium
steel
Description of
Material
Manufacturer
0.20% carbon steel
0.40% carbon steel

39






A
s
-
F
o
r
g
e
d





(
3
)
M
a
c
h
i
n
e
d

a
n
d

P
o
l
i
s
h
e
d











(
4
)
A
1
3
5
3
0
1
2
.
1
1
4
.
2
0
.
8
5
0
.
4
0
0
.
4
7
B
1
4
9
3
3
1
1
.
7
1
4
.
7
0
.
8
0
0
.
3
5
0
.
4
5
A
2
0
5
4
5
1
6
.
5
2
1
.
0
0
.
7
8
0
.
3
7
0
.
4
7
B
1
8
5
4
1
1
2
.
7
2
1
.
6
0
.
5
9
0
.
3
1
0
.
5
3
A
2
7
4
5
9
1
5
.
0
3
2
.
5
0
.
4
6
0
.
2
5
0
.
5
5
B
2
4
0
5
2
1
5
.
8
2
7
.
9
0
.
5
7
0
.
3
1
0
.
5
4
A
3
0
8
6
6
1
8
.
0
3
1
.
5
0
.
5
7
0
.
2
7
0
.
4
8
B
2
7
8
6
0
1
4
.
4
3
1
.
0
0
.
4
6
0
.
2
4
0
.
5
2
R
a
t
i
o










C
o
l
.

(
3
)

C
o
l
.

(
2
)
M
a
t
e
r
i
a
l
3
%

n
i
c
k
e
l

s
t
e
e
l
n
i
c
k
e
l
-
c
h
r
o
m
i
u
m

A
v
e
r
a
g
e

B
r
i
n
e
l
l

H
a
r
d
n
e
s
s

N
u
m
b
e
r











(
1
)
R
a
t
i
o













C
o
l
.

(
4
)

C
o
l
.

(
2
)
M
a
n
u
f
a
c
t
u
r
e
r
0
.
2
0
%

c
a
r
b
o
n

s
t
e
e
l
0
.
4
0
%

c
a
r
b
o
n

s
t
e
e
l
T
e
n
s
i
l
e

S
t
r
e
n
g
t
h
,

T
o
n
s

p
e
r

s
q
.

i
n
.







(
2
)
E
n
d
u
r
a
n
c
e

F
a
t
i
g
u
e

L
i
m
i
t
,

T
o
n
s

p
e
r

s
q
.

i
n
.
R
a
t
i
o












C
o
l
.

(
3
)

C
o
l
.

(
4
)


T
a
b
l
e

2
.
3
:






F
a
t
i
g
u
e

t
e
s
t

r
e
s
u
l
t
s

f
o
r

f
o
u
r

t
y
p
e
s

o
f

c
o
m
m
o
n
l
y

u
s
e
d

f
o
r
g
e
d

s
t
e
e
l
s

w
i
t
h

t
h
e

a
s
-
f
o
r
g
e
d

a
n
d

m
a
c
h
i
n
e
d

a
n
d

p
o
l
i
s
h
e
d




s
u
r
f
a
c
e

c
o
n
d
i
t
i
o
n
s

u
s
e
d

i
n

[
3
]
.



40


Table 2.4: Fatigue test results for a high strength forged steel from two manufacturers
with the as-forged and machined and polished surface conditions used in
[4].




Table 2.5: Relationship between surface carbon content, depth of decarburization,
and fatigue limit for a 605 M36 through-hardened steel [14].





















As-Forged
(2)
Machined and
Polished
(3)
B 118 9.4 49.5 0.19 0.08 0.42
C 138 8.3 49.5 0.17 0.06 0.36
Manufacturer
Tensile
Strength,
Tons per
sq. in.
(1)
Endurance Fatigue Limit,
Tons per sq. in.
Ratio
Col. (2)
Col. (3)
Ratio
Col. (2)
Col. (1)
Ratio
Col. (3)
Col. (1)


41



Table 2.6: Summary of endurance limits and tensile strengths for ground, machined,
hot-rolled, and as-forged surface conditions used to calculate surface
finish factors from [2].




Table 2.7: Values of
R
area parameter used in Murakami equation to predict fatigue
limit for specimens which have surface roughness defects [23].












42

Table 2.8: Results of experimental and predicted fatigue limits using the Murakami
equation for JIS S45C steel specimens having surface roughness defects
[23].





43


Table 2.9: Surface finish modification factors at 10
4
and 10
6
cycles for hardened
machined surfaces calculated using the Goodman, Morrow, and SWT
models [5].



Table 2.10: Constants for use in Eqn. 2.22 for determining surface finish factors for
ground, machined, hot-rolled, and as-forged surfaces [27].



Table 2.11: Fatigue test results on as-forged, as-forged/shot blasted, forged/heat
treated, and forged/ heat treated/ shot blasted specimens of EN15R steel
[1].


Batch
Weibull
Slope

Characteristic Life
(cycles)
AF 4.82 89,809
FS 5.36 140,271
HT 11.27 91,378
HTS 2.04 500,453
AF 5.97 3,820
FS 4.69 3,470
HT 11.44 2,724
AF = as-forged, without heat treatment and shot blasting
HT = forged, heat treated without shot blasting
FS = forged and shot blasted without heat treatment
HTS = forged, heat treated, and shot blasted
/2 = 2200
/2 = 6000

44
Table 2.12: List of typical heat treatment procedures for steels [40].



















Treatment Details Objective
1. Annealing
(a.) Full annealing
(b.) Sheroidising anneal Heat below A
3
.
Spheroidise carbides making
steels amenable to cold forming
processes, and higher carbon
steels suitable for machining.
(c.) Stress relief anneal Heat in range 550-650C.
Relieve internal stresses due to
machining, welding or cold
working.
2. Normalising
Austenitise, (A
3
+30-50C) and
cool in air.
Refine grain size and improve
uniformity of structure.
3. Quenching and Tempering
Austenitise, (A
3
+30-50C),
rapidly cool. Reheat at
intermediate temperature,
e.g. 300-400C.
Obtain best combination of
strength and ductility. By varying
the tempering temperature
possible to obtain desired
properties within a wide range.
4. Austempering
Quench from austenite region
to a temperature suitable for
bainite formation.
Obtain greater ductility while
maintaining high hardness.
Process restricted to small
components.
5. Martempering
Quench from austenite region
to a temperature just above
M
5
, air cool to room
temperature. Temper at
intermediate temperature.
Process gives similar properties
to quenching and tempering,
but greatly reduces possibility of
distortion and quench cracks.
Austenitise, (A
3
+30-50C) and
furnace cool.
Soften hypo-eutectoid steel for
machining or cold working.

45
Table 2.13: Material conditions, test parameters and fatigue lives for notched 4340
low alloy steel specimens with and without decarburization [46].



Table 2.14: Roughness measurements for various types of surface preparations in type
304 stainless steel [47].













Surface
Preparation
Surface Roughness
R, m (rms)
Maximum Depth of
Surface Grooves, m
Ground on silicon
carbide paper
600 grit 0.045 0.18
240 grit 0.48 1.8
50 grit 2.9 8.9
Polished
Mechanically ~0.0075 <0.027
Electrolytically ~0.0075 <0.027

46

(a)

(b)

Figure 2-1: Illustration of (a) closed-die [7] and (b) open-die [8] showing how
material protrudes between upper and lower die.

47



Figure 2-2: The relative shape change of treated and untreated inclusions compared to
shape change of the steel billet for strains of 0.2 and 1.0 at 1000 C

[11].








Figure 2-3: Illustration of a cross section of a forged steel part showing grain flow
[12].



Bulk Material
Inclusions
Treated
Inclusions

48

Figure 2-4: Projection of a defect on to a plane perpendicular to maximum tensile
stress which is used to calculate the square root area parameter [22].




Figure 2-5: Endurance limit versus tensile strength for as-forged surface finish and
polished surface finish specimens [36].

49






(a)

(b)

(c)

50

(d)

Figure 2-6: Brinell hardness versus distance from surface for as-forged specimens
with average core hardness of (a) 185 HB, (b) 200 HB, (c) 260 HB, and
(d) 360 HB [2].










51

(a) (b)


(c) (d)

Figure 2-7: Magnified images (150 x magnification) of a cross section of as-forged
specimens showing surface decarburization with average core hardness of
(a) 185 HB and decarburization ranging from 0.64 mm to 0.76 mm, (b)
200 HB and decarburization ranging from 0.25 mm to 0.30 mm, (c) 260
HB and decarburization ranging from 0.76 mm to 0.89 mm, and (d) 360
HB and decarburization ranging from 0.30 mm to 0.38 mm [2].


52


Figure 2-8: Endurance limit versus tensile strength showing upper and lower limits for
as-forged surface specimens [13].

53

(a)

(b)

(c)

(d)

Figure 2-9: Definition of (a) R
a
, (b) R
z
, (c) R
t
and R
y
, and (d) S
m
to characterize surface
roughness [18].

54


Figure 2-10: Maximum working stress (at 10
7
cycles) versus mean working stress (at
10
7
cycles) for the ground, machined, hot-rolled and as-forged surface
conditions of steel (302 HB-321 HB) [2].



Figure 2-11: Relationship between Brinell hardness and tensile strength for steels [2].

55


Figure 2-12: Plot of endurance limit (defined at 10
7
cycles) versus tensile strength for
steels having ground, machined, hot-rolled and as-forged surface
conditions of steels [2].



Figure 2-13: Surface roughness versus median fatigue life for SAE 1035 and SAE 3130
steels [20].

56


Figure 2-14: Scatter of fatigue test results for SAE 1035 and SAE 3130 steels having a
range of surface roughness [20].


57


Figure 2-15: Configuration and dimensions (mm) of JIS S45C steel fatigue specimens
used in [23].


Figure 2-16: Magnified image of artificially induced surface roughness for annealed
(A) and quenched and tempered (QT) JIS S45C steel with roughness pitch
at 100, 150 and 200 in [23].

58



Figure 2-17: Composite plot of stress versus fatigue life for different types of roughness
in JIS S45C steel specimens [23].




Figure 2-18: Relationship between
R
area /2b and a/2b for annealed (A) and
quenched and tempered (QT) JIS S45C steel fatigue specimens [23].

59


Figure 2-19: Relationship between predicted and experimental fatigue limits for
annealed and quenched and tempered JIS S45C steel [23].




Figure 2-20: Specimen geometry and test set up for the hardened SAE 4140 steel shaft
subjected to bending fatigue in [5].


60

Figure 2-21: Surface finish modification factors for fatigue life ranging from 10
4
to 10
6

cycles for hardened machined surfaces calculated using the Goodman,
Morrow, and SWT models for hardened machined steel surfaces [5].



Figure 2-22: Roughness profile showing definition of the parameter (average radius
of the dominant profile valleys) used in Arola-Ramulu Model [25].


61


Figure 2-23: Fatigue stress concentration factor calculated using Arola-Ramulu and
Neuber models, and experimental fatigue stress concentration factor
versus average surface roughness for a high strength low alloy steel [25].


Figure 2-24: Surface finish modification factor for steels versus tensile strength or
Brinell hardness for ground and polished, machined, hot-rolled, and as-
forged surface conditions [26].

62

Figure 2-25: Stress versus cycles to failure for two type of as-forged surface finish
specimens made from SAE 1035 steel (233 HB-280 HB) and SAE 4063
steel (388 HB-444 HB) [13].



Figure 2-26: S-N curves for steel specimens having machined and polished surface and
as-forged surface conditions.
10
3
10
4
10
5
10
6
10
7
Cycles to Failure, N
f
Strength reduction
due to surface
finish factor by k
s

Forged surface
Polished surface
S
t
r
e
s
s

A
m
p
l
i
t
u
d
e
0.9 S
u


63

(a)


(b)

Figure 2-27: Atomic structure of iron showing (a) body centered cubic, and (b) face
centered cubic [43].



Figure 2-28: Phase diagram for plain carbon steels [43].

64



Figure 2-29: Hardness versus carbon concentration in plain carbon steels [43].



65

Figure 2-30: Tensile strength, yield strength, and reduction in area versus tempering
temperature for 4340 oil quenched steel with martensitic microstructure
[43].




66

(a)

(b)

Figure 2-31: Monotonic and cyclic stress-strain curves for (a) SAE 1045 steel at 595
HB, 500 HB, 450 HB, and 390 HB, as well as (b) SAE 4142 steel at 670
HB, 560 HB, 475 HB, 450 HB, and 380 HB [44].

67



Figure 2-32: Stress amplitude versus plastic strain amplitude for SAE 1045 steel at 595
HB, 500 HB, 450 HB, and 390 HB, and for SAE 4142 steel at 670 HB,
560 HB, 475 HB, 450 HB, and 380 HB [44].

(a) (b)


Figure 2-33: Total strain amplitude versus cycles to failure for strain-controlled
completely-reversed fatigue testing of (a) SAE 1045 steel at 700 HB, 600
HB, and 450 HB, as well as (b) SAE 4142 steel at 670 HB, 560 HB, and
450 HB [44].

68


(a)

(b)

Figure 2-34: Total strain amplitude versus hardness for different cycles to failure for
strain-controlled completely-reversed fatigue testing of (a) SAE 1045
steel, and (b) SAE 4142 steel [44].

69


Figure 2-35: Plot of growth rate (da/dN) versus stress intensity factor range (K) for
martensitic steels having various yield and tensile strengths [45].



Figure 2-36: Specimen configuration and dimensions for a 4340 low alloy steel notched
specimen used to determine the effect of decarburization on fatigue
behavior in [46].

70



Figure 2-37: Magnified image of surface roughness for type 304 stainless steel having
(a) ground and polished surface, and surface roughened with (b) 600 grit,
(c) 240 grit, and (d) 50 grit sand paper [47].


71


Figure 2-38: Crack length versus number of cycles for type 304 stainless steel
specimens at four levels of surface roughness [47].



Figure 2-39: Crack initiation life versus surface roughness for type 304 stainless steel
specimens [47].


72


Figure 2-40: Notch geometry of three-point bending steel specimen with hardness
between 180 HB and 230 HB used for crack growth study in [48].



Figure 2-41: Number of cycles versus average roughness for three-point bending steel
specimens with hardness between 180 HB and 230 HB [48].


















73







Chapter 3





Experimental Procedure and Results


This chapter discusses the specimen configuration and preparation as well as the
test procedures used in this study. There is also a discussion of metallurgical analysis
performed. Results of monotonic tests are discussed and related to the hardness of the
material. Finally, the bending fatigue test results are represented by a series of S-N
curves.

3.1 Material and Specimen Preparation
The material used in this study was a 10B40 steel, which is a common forging
steel. Chemical composition of the material is shown in Table 3.1. The specimens were
hot-forged from normalized rolled bars with diameter of 24 mm and using either gas
furnace heating or induction heating. Forging by gas furnace heating was done at Green
Bay Drop Forge, while forging by induction heating was done at Keystone Forging
Company. Excess material was allowed to protrude out of the space between the upper
and lower dies creating a flash line. The specimens forged using gas furnace heat had the
flash trimmed. However, the specimens forged using induction heat did not have the flash
trimmed. The trimmed flash (gas furnace heated specimens) and untrimmed flash

74
(induction heated specimens) can be seen in Figure 3-1. It should be noted that after the
forging process, specimen hardness was between 85 HRB and 90 HRB. The specimen
configuration and nominal dimensions are shown in Figure 3-2. The specimen diameter
varied slightly from dimensions shown in Figure 3-2, with a deviation of about

0.3
mm.
The specimens were shot cleaned (sand blasted) after the forging process prior to
heat treatment, in order to remove surface scales left from the forging process. It should
be noted, however, that any residual stresses induced during the shot cleaning process
were removed during heat treatment. Specimens forged using induction heat were shot
cleaned by the manufacturer, but the specimens forged by gas furnace heat were shot
cleaned at the University of Toledo Machine Shop. The specimens were heat treated to
four hardness values of 45 HRC, 35 HRC, 25 HRC, and 19 HRC. Heat treatment was
performed using eight samples at a time in a gas furnace at Chrysler Material
Laboratories. In order to evaluate the grain flow effect, some specimens were also
machined from the rolled bar. The diameter of the rolled bar was reduced to about 17 mm
prior to heat treatment by use of a lathe.
Following the heat treatment, the as-forged surface was removed from some
specimens using a CNC machine. The specimen geometry was kept the same, only the
minimum diameter changed from 8 mm to about 6.5 mm by removing 1.5 mm. The
minimum diameter of the machined specimens varied

0.5 mm due to the curvature (i.e.


distortion) of the as-forged part. In addition, the 17 mm diameter bars heat treated for
grain flow testing, were machined to final dimensions (same geometry as as-forged
specimens and with minimum diameter of 7 mm) after heat treatment. All specimens

75
were machined in the Mechanical, Industrial, and Manufacturing Engineering Machine
Shop at the University of Toledo. The machined surface specimens were then polished to
a near mirror surface finish by a polishing machine. Three different grits of aluminum
oxide lapping film with 30 m, 12 m, and 3 m roughness were used for polishing.
Polishing marks coincided with the longitudinal direction of the specimens. The polished
surfaces were carefully examined under magnification to ensure complete removal of
machining marks within the specimen gage section.
Shot cleaning is typically done to forged parts in order to clean the surface, which
in addition may induce beneficial surface compressive residual stresses. To simulate this
effect, following heat treatment, 80 specimens with 10 from each forging method and
each hardness condition were shot cleaned at West Side Sand Blasting. The process used
was representative of a typical process used to clean a forged part of similar size in
industry. A rubber belt tumble blaster, using an S-170 cast steel shot mixture, was used in
the shot cleaning process. All 80 samples were cleaned in the same machine at the same
time. It should be noted that they were exposed to the shots for a total of 10 minutes.
Cantilever and rotating bending fatigue tests were then conducted with the different
groups of specimens. Figure 3-3 shows a flow chart of the different conditions tested.

3.2 Testing Equipment
3.2.1 Monotonic tension and cantilever bending fatigue tests
A closed-loop servo-controlled hydraulic axial load frame in conjunction with a
digital servo-controller was used to conduct the strain-controlled monotonic tension tests
as well as the load-controlled constant amplitude fully-reverse cantilever bending fatigue

76
tests. A 100 KN Instron load cell was used to monitor loads during tension tests. Total
strain was controlled using an extensometer rated as ASTM class B1 [49]. The calibration
of the extensometer was verified using displacement apparatus containing a micrometer
barrel in divisions of 0.0001 in. The extensometer had a gage length of 6 mm and was
capable of measuring strains up to 10%. It should be noted that the specimen had an
hourglass shaped gage section with a difference of 0.08 mm in diameter at the knife edge
of the extensometer and at the minimum cross section. This represents a variation of only
about 1% in diameter.
A 25 KN dynamic Instron load cell was used to control the load during cantilever
bending fatigue tests. A fixture was designed to fix one end of the specimen so that the
specimen major axis was perpendicular to the motion of the loading actuator. Another
fixture was designed to transfer the load from the machine loading actuator to the other
end of the specimen. The design of this fixture incorporated roller bearings to
accommodate any change in the specimen length during cyclic bending and avoiding any
axial stress. Figure 3-4 shows a photo of the test setup used. The test setup was verified
using a specimen with strain gages attached at the maximum stress locations. This can be
seen in Figure 3-5, which shows a plot of calculated strain versus measured strain data as
well as a 45 line, which represents the perfect agreement line. It can be seen that the data
do not deviate from the 45 line. The loads used to verify the fixture ranged from zero to
225 N.




77
3.2.2 Rotating bending fatigue tests
A four-point rotating bending fatigue machine was used to perform rotating
bending fatigue tests. This machine was powered by an electric motor having a constant
speed of 1750 RPM, producing cycling at 29 Hz. The machine was designed to create a
uniform bending moment across the specimen gage section. Weight added to a hook
below the center of the specimen length creates the bending moment. An image of the
four-point bending machine is shown in Figure 3-6. A round strain-gaged bar was used to
verify the bending moment imposed.

3.3 Test Methods and Procedures
3.3.1 Metallurgical analysis
The forging process resulted in surface decarburization and discontinuities
(micro-cracks). In order to determine the extent of decarburization and discontinuities,
samples were prepared for examination under microscope before and after heat treatment
by Chrysler Material Laboratories. Samples were prepared in both the transverse and
longitudinal directions in both the gage and grip sections.
Micro-hardness readings were taken using a Vickers microhardness testing
machine at the University of Toledo. Hardness tests were performed on one sample at
each hardness level for both gas furnace and induction heated specimens. In order to
perform the hardness tests, a cross section of the specimens was taken from the center of
the gage section. Each cross section was first mounted in Bakelite powder. Following the
mounting procedure, each sample was then wet sanded using four different grits of
silicon carbide sanding paper, 240 grit, 320 grit, 400 grit, and 600 grit. After the wet

78
sanding, the samples were then polished on a polishing wheel using a 0.5 alumina
solution followed by a 0.05 alumina solution. Finally the surface was etched using a 0.3%
nital solution. Ten hardness measurements were taken in 0.5 mm increments across the
gage section diameter starting from the surface for each sample. It should be noted that as
a part of the metallurgical analysis performed by Chrysler Material Laboratories, micro-
hardness measurements were taken for some of the test samples. These measurements
were in good agreement with measurements taken at the University of Toledo.

3.3.2 Surface roughness measurements
Surface roughness measurements were made using a Taylor Hobson Tally Surf
probe type profiler, located at the Gear Research Laboratory at Ohio State University.
This machine recorded the profile of the surface which allowed measurement of several
different roughness parameters. Eighteen specimens were measured, each one
representing a different condition, with two measurement made on each specimen. A 10
mm profile of the surface was taken at the center of the grip ends, 90 degrees from the
flash. A 2.5 mm cutoff was used for calculation of roughness parameters. It should be
noted that the gage section could not be used for roughness measurement due to the
hourglass shape.

3.3.3 Monotonic tension tests
Tension tests were performed on four machined and polished specimens, one at
each hardness level. All monotonic tests were performed using test methods specified by
ASTM Standard E8 [50]. In order to protect the extensometer, strain control was used

79
only up to 8% strain, until the point of ultimate tensile strength had been crossed. After
this point, displacement control was used until fracture. INSTRON Bluehill software was
used for the monotonic tension tests. For the elastic and initial yield region (0% to 0.5%
strain) as well as the period up to which the extensometer was removed, a strain rate of
0.0025 mm/mm/min was chosen. This strain rate was three-quarters of the maximum
allowable rate specified by ASTM Standard E8 for the initial yield region. After the
extensometer was removed, a displacement rate of 0.15 mm/min was used.

3.3.4 Cantilever bending fatigue tests
Cantilever bending fatigue tests were performed on specimens from all four
hardness levels, 45 HRC, 35 HRC, 25 HRC, and 19 HRC. Specimens forged using both
gas furnace and induction heating were utilized. Specimens machined and polished from
forgings as well as rolled bars were also subjected to cantilever bending fatigue. In
addition, shot cleaned as-forged parts from both heating conditions were tested. Figure 3-
3 lists the different conditions tested under cantilever bending loading in the form of a
flowchart. The test setup was described in the previous section (refer to Figure 3-4 for
image of the setup). First the specimen was inserted into Apparatus 1 which was used to
transfer the motion of the actuator to the specimen. There was an identifying mark on the
grip end of every specimen from the forging die. The grip end with the mark was inserted
into Apparatus 1 so that the mark was facing upwards. In addition, none of the grip end
was protruding from the face of Apparatus 1. Next, the specimen was fixed to Apparatus
2. The distance from the center of the load train to the minimum diameter of the test
sample (i.e. the moment arm) was kept constant at 64.8 mm for all tests. It should be

80
noted that most specimens were inserted into the fixture so that the forging flash line was
on the specimen neutral surface (i.e. 90 degrees from the location of maximum stress) to
avoid any forging flash/trim effects. However, some testing was also done with the flash
line at the location of maximum stress using the same testing procedures in order to
evaluate the flash effect (see Figure 3-3 for tests performed). Instron SAX software was
used to control the load during the cantilever bending fatigue tests. A mean load of -7.8 N
was used to compensate for the weight of Apparatus 1. Load and displacement means and
amplitudes were automatically recorded throughout each test at regular intervals. The
frequencies used in cantilever bending tests ranged from 1 Hz to 7 Hz, depending on the
load level used. The bending moments used ranged from 65 Nm to 10 Nm resulting in a
life range between 10
4
and 2x10
6
cycles to failure. All tests were conducted using a sine
waveform. Failure was defined by complete fracture of the specimen.

3.3.5 Rotating bending fatigue tests
Rotating bending fatigue tests were performed on specimens forged using gas
furnace heat at the 35 HRC and 19 HRC levels. Both as-forged and polished surface
conditions were tested. It should be noted that the polished specimens tested under
rotating bending loading were machined from the forged specimens (see Figure 3-3 for
specimen conditions and tests performed). Specimens were held in place by manually
tightened tapered collets at both ends. Significant effort was made to ensure only a torque
was applied to the collet nut when tightening, to prevent deformation of the specimen
prior to starting the test. The weight required was applied just as the specimen started to
rotate. Bending moment levels ranging from 27 Nm to 10 Nm were used resulting in a

81
life range between 10
4
and 5x10
6
cycles to failure. In rotating bending fatigue tests,
failure was also defined as complete fracture of the specimen.

3.4 Experimental Results and Analysis
3.4.1 Metallurgical analysis
Photographs of the samples gage section for different conditions are shown in
Figures 3-7 through 3-9. Figure 3-7 shows magnified images of the specimen gage
section prior to heat treatment for both the gas furnace heated and induction heated
conditions. The lighter colored area near the surface represents the decarburized area. It
can be seen that the induction heated specimens had less decarburization, about 160
microns deep, compared to gas furnace heated specimens, about 250 microns deep.
Figure 3-8 shows magnified images of the specimen gage section after heat treatment for
gas forged specimens. Figure 3-9 shows a magnified image of an induction heated
specimen at 45 HRC. Similar to Figure 3-7, the lighter colored area near the surface in
Figures 3-8 and 3-9 represents the decarburized layer. Microhardness measurements
suggest that there is little change in depth of decarburization after heat treatment. In
addition, microstructure consisted of pearlite and ferrite prior to heat treatment and
tempered martensite after heat treatment at all hardness levels.
Grain flow resulting from the forging process can be seen in Figure 3-10. It can be
seen in this figure that the grain flow was fairly uniform with the exception of a more
compressed area at the center where the trim line is located. Figure 3-11 shows an image
of surface irregularities in the as-forged specimens forged by induction heating. It can be
seen in this figure that there are discontinuities as deep as 0.06 mm. For the gas furnace

82
heated specimens discontinuities as deep as 0.1 mm were observed. The discontinuities
are about one third of the decarburization depth for each heating method.
Micro-hardness measurements were also taken on the samples prepared for
microscopy. Figure 3-12 shows a plot of Brinell hardness versus depth below the surface
for gas furnace heated and induction heated forged specimens. The results show that there
is a reduction of hardness in the decarburized area. There is less of a reduction of
hardness for the induction heated specimens, as expected, since the decarburization was
less significant, compared to gas furnace heating.

3.4.2 Surface roughness characterization
Several different surface roughness parameters (R
a
, R
y
, R
t
, R
z
and S
m
) were
determined from the measured surface profiles. These parameters were defined in
Chapter 2 and are repeated here for convenience [18]:
}
=
l
a
dx x y
l
R
0
) (
1
(3.1)
|
.
|

\
|
+ =

= =
n
i
n
i
i i DIN z
v p
n
R
1 1
) (
1
(3.2)
v p t
R R R + = (3.3)

=
=
n
i
i m
S
n
S
1
1
(3.4)
Equation 3.1 represents the average deviation from the mean line over a sampling length.
The parameter R
z
, shown in Eqn. 3.2, is defined as the average summation of the five
highest peaks and the five lowest valleys. In Eqn. 3.3, R
p
is the maximum peak height and
R
v
is the maximum valley depth, both measured from the mean line. The parameter R
y
is

83
the maximum peak to valley height. Equation 3.4 represents the mean spacing between
profile peaks at the mean line. Figures 3-13 and 3-14 show the roughness profiles for
induction heated and gas furnace heated specimens, respectively. Each figure shows a
profile of the as-forged surface prior to heat treatment, as-forged surface at 45 HRC, and
the as-forged and shot-cleaned surface at 45 HRC. It can be seen that there is not much
change in the surface profile after heat treatment and shot cleaning for induction heated
specimens. However, there is some difference observed in the surface profile after shot
cleaning for the gas furnace heated specimens.
Table 3.2 summarizes the roughness measurements. It should be noted that the
values listed in Table 3.2 are an average of two readings with minimal variation between
the two readings. Figure 3-15 shows plots of roughness versus hardness for the roughness
parameters R
a
, R
t
, R
y
, and R
z(DIN)
.

The results show that surface roughness parameter R
a

did not change significantly during heat treatment to achieve different hardness levels.
However, the other roughness parameters, R
t
, R
y
, and R
z(DIN)
, show more variations with
changes in hardness. In addition, the variation in the roughness parameters R
t
, R
y
, and
R
z(DIN)
for as-forged surface specimens was similar to the shot cleaned surface specimens
as hardness changed. It can also be seen that the induction heated specimens had a higher
surface roughness than the gas furnace heated specimens. However, this is believed to be
due to the different shot cleaning process used on the induction heated specimens prior to
heat treatment.




84
3.4.3 Monotonic deformation behavior
The properties determined from monotonic tests were yield strength (YS), ultimate
tensile strength (S
u
), percent elongation (%EL), percent reduction in area (%RA), strength
coefficient (K), and strain hardening exponent (n). A summary of the monotonic
properties for the material is provided in Table 3.3. The monotonic stress-strain curves
are shown in Figure 3-16. As expected, the tensile strength increases as the hardness
increases. Figure 3-17 shows a plot of the tensile strength versus Brinell hardness, along
with a commonly used estimation line [26] based on:
( ) HB 45 . 3 =
u
S (3.5)
where S
u
is in MPa. It can be seen in this figure that Eqn. 3.5 provides a fairly accurate
representation of the data.

3.4.4 Cantilever bending fatigue behavior
Tables 3.4 through 3.7 contain the results from cantilever bending fatigue testing
for 45 HRC, 35 HRC, 25 HRC, and 19 HRC specimens, respectively. These tables list
specimen diameter, heating method, surface condition, bending type loading, test
frequency, load and stress amplitudes, midlife displacement amplitudes, and cycles to
failure. Figure 3-18 shows typical fracture surfaces at different stress amplitudes for each
hardness level tested under cantilever bending loading.
As previously mentioned, the as-forged surface is accompanied by significant
surface roughness and surface discontinuities (see Figure 3-11). As a result, the fatigue
life of the as-forged specimens is dominated by fatigue crack growth. This can be
observed in Figures 3-19 through 3-22 showing superimposed plots of displacement

85
amplitude versus normalized cycles for the different as-forged surface conditions. The
increase in displacement amplitude observed in these figures is due to crack growth and
not cyclic softening due to plastic deformation since the loading was nearly fully elastic.
Crack growth data and analysis for as-forged surface specimens is discussed in Chapter 5.
Figure 3-23 shows plots of displacement amplitude versus cycles to failure for the
machined and polished specimens. The displacement amplitude observed in Figure 3-23
remains constant for a majority of the fatigue life, implying that fatigue life is dominated
by fatigue crack initiation, as would be expected for specimens with a polished surface. It
should be noted that only one test from each stress level was included in these figures,
although multiple tests were run at each stress level.
The flash line effect was investigated during some cantilever bending fatigue tests
by positioning the flash line at the maximum stress location. The effect was found to be
insignificant, as will be discussed in Chapter 4.
The fatigue test results are represented in a series of stress-life diagrams shown in
Figures 3-24 through 3-28, with one figure for each surface condition. The stress
amplitudes applied to the tests specimens were determined from the elastic bending
equation:
3
32
d
M
I
c M
a a
a
t
o = = (3.6)
Figure 3-29 shows plots of midlife load versus displacement for tests with the highest
stress amplitudes used in the S-N line fits for induction heated and polished specimens.
This figure shows mostly linear behavior and supports the use of Eqn. 3.6. Fatigue
strength for a given number of cycles to failure can be determined from the S-N line
equation represented by:

86
( )
B
f a
N A =
A
=
2
o
o (3.7)
When performing the least squares fit to determine the intercept A and slope B, the stress
amplitude (/2) was the independent variable and the cycles to failure (N
f
) was the
dependent variable. To generate the A and B values, all data in the stress-life figures were
used, with the exception of data from test levels where run-out (i.e. N
f
> 10
6
cycles) had
occurred. The highest stress amplitudes (S
a
> 1000 MPa) for some of the 45 HRC
specimens were not used in the fits either due to the possibility of surface plastic
deformation.

3.4.5 Rotating bending fatigue behavior
Tables 3.5 and 3.7 also contain the results from rotating bending fatigue testing
for 35 HRC and 19 HRC specimens, respectively, as previously discussed. Figure 3-30
shows typical fracture surfaces at different stress amplitudes for each hardness level
tested under rotating bending loading. Equation 3.6 was also used to calculate the stress
amplitudes in the rotating bending fatigue tests. Fatigue strength for a given number of
cycles to failure for rotating bending tests was also determined from Eqn. 3.7. Similar to
the cantilever bending tests, the A and B values were generated using all data in the
stress-life figures with the exception of data from tests levels where run-out (N
f
> 10
6

cycles) had occurred.
As mentioned earlier, in some of the cantilever bending tests the effect of the
flash line was investigated by positioning the flash line at the maximum stress location
and the effect was found to be insignificant. Similarly, in rotating bending tests of the as-
forged surface finish specimens, cracks did not initiate at the flash line for about half of

87
the tests. In addition, the crack initiation location was independent of the fatigue life, as
the specimens with crack initiating at the flash did not always fail sooner than the
specimens with cracks not initiating at the flash line. Therefore, it can be concluded that
the flash had little or no effect under rotating bending conditions of the as-forged surface
finish specimens.
Figure 3-31 shows plots of stress amplitude versus cycles to failure for rotating
bending fatigue tests. The rotating bending fatigue test results show similar results to
cantilever bending fatigue tests, as will be discussed in Chapter 4.
Table 3.8 lists values A and B in Equation 3.7 obtained from best line fits to the
experimental data for each test condition for both cantilever and rotating bending fatigue
tests. This table also lists the fatigue limits (N
f
2x10
6
) and fatigue strength at 10
6
cycles.
The two values are different for some test conditions.












88
Table 3.1: Material composition of 10B40 steel (Courtesy of Robert Cryderman of
Gerdau-MacSteel).




Table 3.2: Summary of surface roughness measurements (m) of as-forged, as-
forged and heat treated, and as-forged and heat treated and shot cleaned
for 10B40 steel.



Hardness
(HRC)
Surface
Condition
R
a
R
t
R
y
R
z
DIN S
m
Induction 7.7 52.2 52.2 41.0 298.2
Gas furnace 6.3 36.9 35.5 30.3 275.6
Induction 9.5 40.2 38.7 34.7 811.8
Induction-
shot cleaned
8.1 50.0 47.4 42.4 306.1
Gas furnace 6.5 31.9 31.7 26.4 630.1
Gas furnace-
shot cleaned
4.8 33.5 29.1 28.0 240.9
Induction 7.5 46.2 44.7 40.6 270.5
Induction-
shot cleaned
8.3 49.4 48.4 40.7 385.9
Gas furnace 6.9 45.3 40.5 36.5 233.2
Gas furnace-
shot cleaned
5.6 47.8 45.7 36.8 195.5
Induction 8.6 55.6 51.7 47.0 347.0
Induction-
shot cleaned
8.9 50.3 50.3 40.1 508.5
Gas furnace 8.1 51.2 50.9 43.9 315.7
Gas furnace-
shot cleaned
5.1 36.1 34.4 31.3 271.5
Induction 8.3 54.0 49.4 42.8 336.6
Induction-
shot cleaned
8.0 49.6 49.0 41.4 290.6
Gas furnace 5.3 36.5 34.5 29.8 291.1
Gas furnace-
shot cleaned
6.6 51.4 49.9 39.0 274.6
* Values are an average of 2 measurements
19
Prior to
heat treat
45
35
25
Element % Weight Element % Weight
C 0.38 Sn 0.012
Mn 0.94 Al 0.03
P 0.01 V 0.007
S 0.021 B 0.0017
Si 0.24 Ti 0.042
Ni 0.08 Nb 0.002
Cr 0.16 N 0.0084
Mo 0.03
Cu 0.2
Fe Balance

89

Table 3.3: Summary of monotonic tension test results.





































Properties 45 HRC 35 HRC 25 HRC 19 HRC
HB 421 327 253 220
S
y
(MPa) 1375 1034 763 615
S
u
(MPa) 1524 1110 852 701
%EL 32.5 39.8 54.2 58.8
%RA 41.8 51.0 55.6 59.9
K (MPa) 1865 1343 1051 883
n 0.047 0.046 0.061 0.078

90
Table 3.4: Summary of fatigue test results with different heating methods and surface
finish conditions for 45 HRC specimens.


[a] Max stress is at flash.
[b] Test stopped due to pump failure.
Load
Amplitude
P
a
,
(N)
Stress
Amplitude

a
,
(Mpa)

a
(mm)
Top Bottom
Induction As-Forged IF411 Cantilever 8.10 1 995 1237 2.74 3,045 2.62 1.03
Induction As-Forged IF41 Cantilever 8.08 2 763 955 2.09 10,527 3.15 0.91
Induction As-Forged IF42 Cantilever 8.19 2 772 927 2.06 12,595 1.06 3.22
Induction As-Forged IF43 Cantilever 8.12 3 691 851 1.83 15,586 3.87 2.14
Induction As-Forged IF44 Cantilever 7.97 2 643 838 1.85 15,954 3.42 3.15
Induction As-Forged IF46 Cantilever 8.12 3 500 617 1.29 36,021 4.94 0.74
Induction As-Forged IF45 Cantilever 8.15 2 340 415 0.86 123,377 4.65 -
Induction As-Forged IF47 Cantilever 8.14 2 340 416 0.88 127,643 3.92 0.75
Induction As-Forged IF48 Cantilever 8.12 3 211 260 0.60 536,056 5.67 -
Induction As-Forged IF410 Cantilever 8.10 5 208 258 0.54 603,011 5.68 -
Induction As-Forged IF412 Cantilever 8.13 5 187 230 0.52 >2,000,000 - -
Induction As-Forged IF49 Cantilever 8.23 7 175 207 0.39 >2,000,000 - -
Induction Machined IM47 Cantilever 6.78 1 567 1200 3.07 10,033 2.63 3.19
Induction Machined IM43 Cantilever 7.15 2 554 1000 2.10 25,203 3.07 3.28
Induction Machined IM41 Cantilever 7.46 3 567 901 1.95 30,663 3.65 -
Induction Machined IM46 Cantilever 6.49 2 373 899 2.23 75,663 3.29 -
Induction Machined IM45 Cantilever 7.21 3 454 799 1.80 >1,200,000 - -
Induction Machined IM44 Cantilever 7.22 3 399 700 1.50 >1,400,000 - -
Induction Shot-Cleaned IS44 Cantilever 7.98 2 732 950 2.17 10,569 3.51 2.70
Induction Shot-Cleaned IS45 Cantilever 8.04 5 747 950 2.20 11,445 3.54 0.95
Induction Shot-Cleaned IS46 Cantilever 8.22 5 588 700 1.65 25,624 4.22 1.02
Induction Shot-Cleaned IS47 Cantilever 8.06 4 474 599 1.42 43,587 3.67 0.89
Induction Shot-Cleaned IS410 Cantilever 8.00 4 466 601 1.32 45,142 3.62 1.45
Induction Shot-Cleaned IS48 Cantilever 7.99 4 271 350 0.76 398,727 5.13 -
Induction Shot-Cleaned IS49 Cantilever 8.06 4 277 350 0.79 437,028 - 5.28
Induction Shot-Cleaned IS412 Cantilever 8.17 5 214 270 0.68 >2,000,000 - -
Gas As-Forged GF44 Cantilever 7.89 2 688 924 2.02 10,470 3.81 2.44
Gas As-Forged GF47 Cantilever 7.96 2 709 927 2.16 10,700 3.28 1.81
Gas As-Forged GF42 Cantilever 7.90 3 473 635 1.36 44,301 1.96 3.82
Gas As-Forged GF45 Cantilever 7.89 3 478 643 1.39 45,513 3.81 3.12
Gas As-Forged GF43 Cantilever 7.93 5 316 419 0.90 111,609 0.75 4.54
Gas As-Forged GF41 Cantilever 7.86 3 314 428 0.89 160,274 3.09 3.87
Gas As-Forged GF414 Cantilever 7.88 5 223 301 0.70 471,236 5.24 -
Gas As-Forged GF48 Cantilever 7.98 7 199 259 0.55 >1,800,000 - -
Gas As-Forged GF46 Cantilever 7.99 7 196 253 0.56 >2,000,000 - -
Gas As-Forged GF412
[a]
Cantilever 7.86 2 662 900 2.11 13,327 3.79 0.75
Gas As-Forged GF413
[a]
Cantilever 7.80 3 439 610 1.33 57,910 5.08 2.45
Gas As-Forged GF410
[b]
Cantilever 8.06 3 448 586 1.48 >24,113 - -
Gas As-Forged GF411
[a]
Cantilever 7.78 3 436 586 1.34 51,969 3.59 4.61
Gas As-Forged GF49
[a]
Cantilever 7.80 5 287 377 0.95 203,432 1.72 5.21
Gas Machined GM45 Cantilever 6.72 1 552 1199 2.78 10,629 3.11 1.94
Gas Machined GM47 Cantilever 6.81 2 479 1001 2.27 24,000 2.47 -
Gas Machined GM43 Cantilever 7.23 2 573 1001 2.21 25,996 2.50 3.67
Gas Machined GM42 Cantilever 7.23 3 515 899 1.99 45,237 1.12 3.23
Gas Machined GM44 Cantilever 7.16 3 473 850 1.84 64,780 3.51 1.89
Gas Machined GM48 Cantilever 6.95 2 433 851 1.96 73,393 - 3.03
Gas Machined GM46 Cantilever 6.70 3 365 800 1.81 >1,200,000 - -
Gas Machined GM41 Cantilever 7.10 3 330 609 1.36 >1,000,000 - -
Gas Shot-Cleaned GS42 Cantilever 7.84 2 655 899 2.10 12,535 3.18 1.00
Gas Shot-Cleaned GS43 Cantilever 7.81 2 648 900 2.04 14,956 3.64 3.47
Gas Shot-Cleaned GS44 Cantilever 7.85 2 512 699 1.58 32,906 3.77 2.81
Gas Shot-Cleaned GS45 Cantilever 7.82 3 435 600 1.31 45,553 - 3.81
Gas Shot-Cleaned GS47 Cantilever 7.85 4 366 500 1.14 123,386 4.02 -
Gas Shot-Cleaned GS46 Cantilever 7.87 4 369 501 1.15 215,330 - 3.94
Gas Shot-Cleaned GS48 Cantilever 8.04 5 231 315 0.80 >2,000,000 - -
Rolled Bar Machined R42 Cantilever 6.90 2 597 1199 3.18 9,625 1.67 3.30
Rolled Bar Machined R43 Cantilever 6.89 2 594 1199 3.15 11,508 3.52 2.38
Rolled Bar Machined R41 Cantilever 6.82 2 433 900 2.35 61,736 - 3.61
Rolled Bar Machined R44 Cantilever 6.90 2 448 899 2.36 55,019 3.65 -
Rolled Bar Machined R45 Cantilever 6.96 4 409 800 1.79 >2,000,000 - -
N
f
Crack Location
and Depth
Bending
Loading
Diameter
D
0,
(mm)
Test
Freq.,
(Hz)
Midlife
Heating
Surface
Condition
Specimen
ID

91
Table 3.5: Summary of fatigue test results with different heating methods and surface
finish conditions for 35 HRC specimens.



Load
Amplitude
P
a
,
(N)
Stress
Amplitude

a
,
(Mpa)

a
(mm)
Top Bottom
Induction As-Forged IF34 Cantilever 8.15 2 789 961 2.09 4,568 3.64 2.84
Induction As-Forged IF36 Cantilever 7.98 2 713 928 1.98 5,948 3.50 3.14
Induction As-Forged IF35 Cantilever 8.10 3 510 633 1.37 37,548 4.58 2.97
Induction As-Forged IF31 Cantilever 8.13 2 515 634 1.36 41,560 1.75 4.90
Induction As-Forged IF32 Cantilever 7.94 3 316 416 0.91 193,529 4.44 3.27
Induction As-Forged IF33 Cantilever 8.00 4 318 410 0.85 224,433 2.10 4.97
Induction As-Forged IF38 Cantilever 8.16 4 268 325 0.69 679,675 - 5.23
Induction As-Forged IF37 Cantilever 8.21 5 251 303 0.66 >1,800,000 - -
Induction Machined IM38 Cantilever 6.46 2 368 899 2.30 15,169 3.01 1.19
Induction Machined IM31 Cantilever 6.66 2 403 901 2.10 19,075 3.60 1.91
Induction Machined IM32 Cantilever 6.39 3 336 849 1.95 27,056 3.29 1.96
Induction Machined IM36 Cantilever 6.72 2 369 801 1.94 31,041 3.68 0.70
Induction Machined IM37 Cantilever 6.57 2 344 801 1.97 32,650 2.72 2.84
Induction Machined IM33 Cantilever 6.47 3 287 699 1.67 81,521 1.18 3.50
Induction Machined IM35 Cantilever 6.47 3 286 698 1.81 97,907 - 3.27
Induction Machined IM310 Cantilever 6.78 4 302 640 1.48 >2,000,000 - -
Induction Machined IM39 Cantilever 6.64 3 267 602 1.41 >1,200,000 - -
Induction Shot-Cleaned IS33 Cantilever 8.22 2 715 850 1.89 7,431 3.65 2.50
Induction Shot-Cleaned IS32 Cantilever 8.12 2 689 850 1.91 7,488 3.36 3.12
Induction Shot-Cleaned IS35 Cantilever 8.02 3 492 631 1.39 36,335 - 3.98
Induction Shot-Cleaned IS36 Cantilever 8.03 3 494 630 1.45 40,818 3.44 2.95
Induction Shot-Cleaned IS34 Cantilever 8.06 4 396 500 1.14 131,591 4.00 -
Induction Shot-Cleaned IS31 Cantilever 8.04 4 393 499 1.12 192,088 - 3.85
Induction Shot-Cleaned IS37 Cantilever 8.10 5 290 365 1.05 >2,000,000 - -
Gas As-Forged GF36 Cantilever 7.85 3 658 900 2.05 5,954 3.11 3.20
Gas As-Forged GF31 Cantilever 7.89 2 453 610 1.41 24,362 3.83 1.92
Gas As-Forged GF34 Cantilever 7.84 3 436 599 1.35 41,008 3.58 3.29
Gas As-Forged GF33 Cantilever 7.90 3 298 399 0.86 142,974 1.38 4.25
Gas As-Forged GF35 Cantilever 7.85 3 293 400 0.88 196,739 3.72 2.13
Gas As-Forged GF32 Cantilever 7.89 4 242 326 0.71 326,880 2.08 4.34
Gas As-Forged GF37 Cantilever 7.87 3 240 326 0.71 361,997 2.46 2.99
Gas As-Forged GF38 Cantilever 7.93 5 208 275 0.60 732,894 5.00 -
Gas As-Forged GF321 Cantilever 7.96 5 191 250 0.58 626,864 - 4.98
Gas As-Forged GF310 Cantilever 7.92 3 189 251 0.57 761,631 5.22 -
Gas As-Forged GF322 Cantilever 7.88 5 171 231 0.49 1,326,413 4.19 4.03
Gas As-Forged GF323 Cantilever 7.86 5 171 232 0.50 >2000000 - -
Gas As-Forged GF311 Rotating 8.03 29 - 551 - 43,776 - -
Gas As-Forged GF312 Rotating 8.03 29 - 546 - 44,806 - -
Gas As-Forged GF313 Rotating 8.00 29 - 399 - 96,294 - -
Gas As-Forged GF314 Rotating 7.89 29 - 394 - 112,877 - -
Gas As-Forged GF316 Rotating 8.00 29 - 275 - 701,818 - -
Gas As-Forged GF315 Rotating 7.98 29 - 276 - 823,994 - -
Gas As-Forged GF318 Rotating 8.03 29 - 229 - 850,611 - -
Gas As-Forged GF317 Rotating 7.96 29 - 232 - 993,485 - -
Gas As-Forged GF319 Rotating 8.00 29 - 231 - >4,817,420 - -
Gas Machined GM31 Rotating 6.88 29 - 805 - 23,885 - -
Gas Machined GM32 Rotating 6.88 29 - 743 - 61,901 - -
Gas Machined GM35 Rotating 6.87 29 - 746 - 66,694 - -
Gas Machined GM33 Rotating 6.82 29 - 698 - 135,213 - -
Gas Machined GM34 Rotating 6.85 29 - 651 - 190,348 - -
Gas Machined GM36 Rotating 6.84 29 - 653 - 316,646 - -
Gas Machined GM38 Rotating 6.95 29 - 615 - 352,256 - -
Gas Machined GM310 Rotating 6.98 29 - 577 - >2,400,000 - -
Gas Shot-Cleaned GS33 Cantilever 7.90 2 597 800 1.81 8,824 2.96 2.40
Gas Shot-Cleaned GS32 Cantilever 7.94 2 606 800 1.86 11,622 3.40 2.31
Gas Shot-Cleaned GS37 Cantilever 7.91 3 450 600 1.39 34,712 3.78 2.45
Gas Shot-Cleaned GS35 Cantilever 7.93 3 448 593 1.33 49,752 2.50 3.80
Gas Shot-Cleaned GS31 Cantilever 7.88 4 333 449 1.01 100,596 4.41 0.87
Gas Shot-Cleaned GS36 Cantilever 7.86 4 294 399 0.86 232,621 1.68 4.25
Gas Shot-Cleaned GS34 Cantilever 7.90 4 300 402 0.88 261,255 4.45 1.90
Gas Shot-Cleaned GS38 Cantilever 7.82 5 210 290 0.82 >2,000,000 - -
N
f
Crack Location
and Depth
Midlife
Bending
Loading
Diameter
D
0,
(mm)
Test
Freq.,
(Hz)
Heating
Surface
Condition
Specimen
ID

92
Table 3.6: Summary of fatigue test results with different heating methods and surface
finish conditions for 25 HRC specimens.


[a] Max stress is at flash.


Load
Amplitude
P
a
,
(N)
Stress
Amplitude

a
,
(Mpa)

a
(mm)
Top Bottom
Induction As-Forged IF24 Cantilever 8.11 2 565 700 1.65 5,653 2.78 2.82
Induction As-Forged IF21 Cantilever 7.89 2 453 610 1.34 14,644 3.52 3.00
Induction As-Forged IF27 Cantilever 7.97 2 468 610 1.39 15,654 3.56 3.54
Induction As-Forged IF26 Cantilever 7.86 2 367 500 1.14 48,725 3.32 3.78
Induction As-Forged IF22 Cantilever 8.09 2 401 500 1.05 53,546 4.94 2.20
Induction As-Forged IF23 Cantilever 8.04 3 315 400 0.86 149,703 2.65 5.11
Induction As-Forged IF25 Cantilever 7.94 2 302 399 0.84 160,256 4.39 1.64
Induction As-Forged IF28 Cantilever 8.05 6 257 325 0.69 339,316 0.13 4.35
Induction As-Forged IF29 Cantilever 8.02 6 235 302 0.63 829,847 4.58 0.76
Induction As-Forged IF210 Cantilever 8.15 5 230 280 0.61 >2,000,000 - -
Induction Machined IM24 Cantilever 6.23 3 220 600 1.72 26,755 - 2.41
Induction Machined IM28 Cantilever 6.21 2 218 601 1.60 49,524 2.62 1.76
Induction Machined IM21 Cantilever 6.20 3 216 599 1.54 58,486 3.05 2.61
Induction Machined IM22 Cantilever 6.14 3 193 550 1.41 151,180 2.72 -
Induction Machined IM23 Cantilever 6.22 3 200 550 1.29 181,314 3.11 -
Induction Machined IM26 Cantilever 6.30 4 199 524 1.29 443,419 - 2.99
Induction Machined IM25 Cantilever 5.70 4 141 502 1.30 >2,000,000 - -
Induction Shot-Cleaned IS27 Cantilever 8.05 3 474 600 1.36 15,540 3.07 2.21
Induction Shot-Cleaned IS24 Cantilever 7.95 3 458 602 1.35 17,806 3.21 3.55
Induction Shot-Cleaned IS29 Cantilever 8.13 3 406 499 1.37 53,044 2.72 4.28
Induction Shot-Cleaned IS28 Cantilever 8.16 3 413 501 1.11 54,327 3.91 2.91
Induction Shot-Cleaned IS26 Cantilever 8.13 4 325 400 0.90 189,867 4.70 1.10
Induction Shot-Cleaned IS25 Cantilever 8.06 4 316 399 0.90 195,899 4.13 1.66
Induction Shot-Cleaned IS210 Cantilever 8.14 5 249 305 0.80 >2,000,000 - -
Gas As-Forged GF22 Cantilever 7.92 2 460 611 1.36 18,141 3.00 2.98
Gas As-Forged GF24 Cantilever 7.87 2 449 609 1.33 19,456 3.14 3.44
Gas As-Forged GF26 Cantilever 7.85 3 366 499 1.11 47,540 3.81 2.49
Gas As-Forged GF21 Cantilever 7.95 3 381 500 1.18 51,475 3.56 2.18
Gas As-Forged GF25 Cantilever 8.05 3 317 400 0.90 143,421 2.08 4.02
Gas As-Forged GF23 Cantilever 7.84 3 292 400 0.88 147,839 2.29 3.67
Gas As-Forged GF27 Cantilever 7.97 4 229 298 0.67 401,789 1.05 4.46
Gas As-Forged GF29 Cantilever 7.93 5 210 278 0.62 667,749 1.59 5.02
Gas As-Forged GF28 Cantilever 7.94 5 213 281 0.60 843,878 5.13 1.74
Gas As-Forged GF217 Cantilever 7.89 3 194 260 0.56 733,004 5.01 -
Gas As-Forged GF218 Cantilever 7.90 5 179 240 0.51 1,220,147 1.50 4.90
Gas As-Forged GF226 Cantilever 7.86 3 176 240 0.50 1,495,910 5.16 -
Gas As-Forged GF227 Cantilever 7.91 5 169 225 0.46 >2,000,000 - -
Gas As-Forged GF221
[a]
Cantilever 7.87 5 443 600 1.45 16,011 3.34 3.20
Gas As-Forged GF222
[a]
Cantilever 7.90 3 299 400 0.93 131,242 4.78 2.22
Gas As-Forged GF223
[a]
Cantilever 7.85 3 294 401 0.90 156,151 3.02 4.50
Gas As-Forged GF224
[a]
Cantilever 7.91 5 224 300 0.69 604,423 - 5.38
Gas Shot-Cleaned GS21 Cantilever 7.83 3 436 599 1.37 14,608 3.29 2.99
Gas Shot-Cleaned GS22 Cantilever 7.93 3 453 601 1.35 16,476 3.54 3.66
Gas Shot-Cleaned GS24 Cantilever 7.93 3 378 501 1.12 46,509 3.03 3.80
Gas Shot-Cleaned GS23 Cantilever 7.88 3 371 501 1.13 49,073 4.11 2.12
Gas Shot-Cleaned GS25 Cantilever 7.87 4 296 400 0.89 146,307 4.12 1.06
Gas Shot-Cleaned GS26 Cantilever 7.91 4 299 400 0.88 164,362 4.29 2.91
Gas Shot-Cleaned GS27 Cantilever 8.00 3 216 290 0.78 >2,000,000 - -
Rolled bar Machined R21 Cantilever 6.97 2 308 600 1.51 65,147 3.61 1.19
Rolled bar Machined R22 Cantilever 6.82 2 288 599 1.49 79,479 3.29 2.16
Rolled bar Machined R23 Cantilever 6.97 2 282 549 1.39 184,846 3.80 1.85
Rolled bar Machined R25 Cantilever 6.95 2 267 525 1.31 268,795 3.60 -
Rolled bar Machined R24 Cantilever 6.97 3 269 525 1.31 462,659 3.82 -
Rolled bar Machined R26 Cantilever 6.96 3 256 501 1.22 >2,000,000 - -
Heating
Surface
Condition
Specimen
ID
N
f
Crack Location
and Depth
Midlife
Bending
Loading
Diameter
D
0,
(mm)
Test
Freq.,
(Hz)

93
Table 3.7: Summary of fatigue test results with different heating methods and surface
finish conditions for 19 HRC specimens.


[a] Test stopped due to machine failure.
[b] Test invalid due to excess vibration.
Load
Amplitude
P
a
,
(N)
Stress
Amplitude

a
,
(Mpa)

a
(mm)
Top Bottom
Induction As-Forged IF17 Cantilever 8.12 3 404 499 1.08 21,938 3.16 3.87
Induction As-Forged IF11 Cantilever 8.14 3 409 500 1.10 22,916 2.46 2.99
Induction As-Forged IF12 Cantilever 8.37 3 356 400 0.84 70,546 4.01 2.09
Induction As-Forged IF15 Cantilever 8.20 3 335 400 0.86 89,766 4.24 3.27
Induction As-Forged IF13 Cantilever 8.12 4 242 299 0.62 478,716 4.19 1.27
Induction As-Forged IF14 Cantilever 8.03 4 235 300 0.63 623,066 4.63 1.14
Induction As-Forged IF16 Cantilever 8.14 5 225 276 0.61 617,962 4.53 -
Induction As-Forged IF19 Cantilever 8.07 5 208 261 0.56 1,263,112 4.95 -
Induction As-Forged IF18
[a]
Cantilever 8.12 5 - 260 - >101,200 - -
Induction As-Forged IF110 Cantilever 7.96 5 190 249 0.51 >2,000,000 - -
Induction Machined IM11 Cantilever 6.80 3 262 551 1.44 29,458 3.21 2.93
Induction Machined IM15 Cantilever 6.00 3 181 552 1.46 29,602 1.43 2.65
Induction Machined IM16 Cantilever 6.23 3 182 498 1.33 76,764 2.42 3.03
Induction Machined IM17 Cantilever 6.52 3 210 500 1.27 84,222 2.59 1.58
Induction Machined IM110 Cantilever 5.67 3 131 475 1.22 187,377 - 4.00
Induction Machined IM19 Cantilever 6.74 4 220 475 1.21 200,514 2.98 -
Induction Machined IM14 Cantilever 6.84 4 218 450 1.17 87,404 2.79 1.13
Induction Machined IM12 Cantilever 5.75 4 130 450 1.17 399,170 2.76 -
Induction Machined IM18 Cantilever 5.95 5 144 451 1.12 >2,000,000 - -
Induction Machined IM13 Cantilever 6.21 4 144 398 1.01 >2,000,000 - -
Induction Shot-Cleaned IS11 Cantilever 8.11 3 402 498 1.10 24,016 2.52 4.09
Induction Shot-Cleaned IS12 Cantilever 7.90 3 374 500 1.12 33,877 3.46 3.62
Induction Shot-Cleaned IS14 Cantilever 8.00 4 311 401 0.89 127,446 4.06 3.75
Induction Shot-Cleaned IS13 Cantilever 7.86 4 294 400 0.85 163,763 3.42 3.67
Induction Shot-Cleaned IS16 Cantilever 7.89 5 261 351 0.76 335,930 4.02 1.27
Induction Shot-Cleaned IS15 Cantilever 8.03 5 274 350 0.77 437,011 4.13 0.60
Induction Shot-Cleaned IS18 Cantilever 7.98 5 243 315 0.70 >2,000,000 - -
Gas As-Forged GF11 Cantilever 7.93 3 378 501 1.11 20,884 3.00 3.01
Gas As-Forged GF12 Cantilever 7.92 3 376 500 1.11 23,272 3.71 3.06
Gas As-Forged GF14 Cantilever 7.94 3 303 400 0.93 62,069 3.74 2.19
Gas As-Forged GF13 Cantilever 7.90 3 299 400 0.92 69,427 2.27 3.76
Gas As-Forged GF16 Cantilever 7.90 5 223 299 0.66 332,199 4.43 1.41
Gas As-Forged GF15 Cantilever 7.86 5 220 300 0.66 426,619 4.50 2.42
Gas As-Forged GF17 Cantilever 7.87 5 184 249 0.58 2,027,745 4.75 -
Gas As-Forged GF18 Cantilever 7.88 5 167 225 0.48 >2,500,000 - -
Gas As-Forged GF112 Rotating 7.81 29 - 508 - 17,274 - -
Gas As-Forged GF111 Rotating 7.84 29 - 502 - 18,003 - -
Gas As-Forged GF114 Rotating 7.85 29 - 396 - 74,669 - -
Gas As-Forged GF113 Rotating 7.81 29 - 397 - 85,245 - -
Gas As-Forged GF116 Rotating 7.82 29 - 299 - 260,109 - -
Gas As-Forged GF115 Rotating 7.93 29 - 301 - 323,123 - -
Gas As-Forged GF117
[b]
Rotating 7.90 29 - 249 - 250,061 - -
Gas As-Forged GF118
[b]
Rotating 7.90 29 - 222 - 783,155 - -
Gas As-Forged GF119 Rotating 7.85 29 - 227 - 1,192,410 - -
Gas As-Forged GF121 Rotating 7.89 29 - 223 - >2,500,000 - -
Gas As-Forged GF120 Rotating 7.83 29 - 207 - >2,344,979 - -
Gas Machined GM18 Rotating 6.87 29 - 557 - 26,323 - -
Gas Machined GM19 Rotating 6.97 29 - 549 - 27,846 - -
Gas Machined GM12 Rotating 6.91 29 - 494 - 87,674 - -
Gas Machined GM13 Rotating 6.86 29 - 497 - 103,078 - -
Gas Machined GM16 Rotating 6.89 29 - 452 - 116,403 - -
Gas Machined GM14 Rotating 6.83 29 - 448 - 195,718 - -
Gas Machined GM17 Rotating 7.00 29 - 453 - 239,866 - -
Gas Machined GM15 Rotating 6.80 29 - 397 - >3,000,000 - -
Gas Shot-Cleaned GS11 Cantilever 7.94 3 379 500 1.16 18,154 3.72 2.81
Gas Shot-Cleaned GS12 Cantilever 7.82 3 361 499 1.12 23,564 3.44 2.14
Gas Shot-Cleaned GS14 Cantilever 7.86 4 295 401 0.91 93,147 1.92 3.54
Gas Shot-Cleaned GS13 Cantilever 7.90 4 298 400 0.88 121,516 3.90 2.37
Gas Shot-Cleaned GS16 Cantilever 7.83 5 243 334 0.76 244,863 1.57 4.41
Gas Shot-Cleaned GS15 Cantilever 7.86 5 248 336 0.71 364,347 1.59 4.23
Gas Shot-Cleaned GS18 Cantilever 7.84 5 208 285 0.65 >2,000,000 - -
Crack Location
and Depth
N
f
Midlife
Test
Freq.,
(Hz)
Heating
Diameter
D
0,
(mm)
Bending
Loading
Specimen
ID
Surface
Condition

94
Table 3.8: S-N line equation constants and fatigue strengths.














Hardness
(HRC)
Surface
Condition
Bending
Type
A (MPa) B
S
e
(MPa)
at 10
6
Cycles
S
e
(MPa)
at S-N
Line Knee
Life at S-N
Line Knee
Machined Cantilever 5137 -0.162 800 800 99,188
Induction Cantilever 20697 -0.332 231 231 774,788
Induction-shot
cleaned
Cantilever 11454 -0.272 325 270 987,037
Gas furnace Cantilever 16025 -0.305 270 258 749,371
Gas furnace-
shot cleaned
Cantilever 8960 -0.243 315 315 940,464
Cantilever 4041 -0.154 640 640 152,294
Rotating 2232 -0.100 577 577 794,771
Induction Cantilever 6492 -0.223 303 303 911,309
Induction-shot
cleaned
Cantilever 3992 -0.174 365 365 938,089
Cantilever 9356 -0.265 241 230 1,193,604
Rotating 6937 -0.241 248 231 1,353,191
Gas furnace-
shot cleaned
Cantilever 6373 -0.224 290 290 952,269
Machined Cantilever 1131 -0.060 501 501 861,211
Induction Cantilever 3764 -0.188 281 280 1,014,726
Induction-shot
cleaned
Cantilever 3047 -0.167 305 305 998,110
Gas furnace Cantilever 5519 -0.223 255 225 1,756,732
Gas furnace-
shot cleaned
Cantilever 3313 -0.176 290 290 987,513
Cantilever 1276 -0.082 451 451 326,286
Rotating 1727 -0.111 397 397 590,398
Induction Cantilever 2584 -0.164 266 249 1,509,083
Induction-shot
cleaned
Cantilever 1926 -0.132 315 315 924,801
Cantilever 3015 -0.181 249 225 1,748,178
Rotating 3180 -0.187 241 225 1,445,652
Gas furnace-
shot cleaned
Cantilever 2201 -0.149 285 285 943,895
45
35
25
19
Machined
Gas furnace
Machined
Gas furnace

95



(a) (b) (c)

Figure 3-1: Photograph of (a) gas furnace heated specimen, (b) induction heated
specimen, and (c) machined and polished specimen.







96

R83
8
16
178
64 51

(a)

8
16
102
31 40

(b)

Figure 3-2: Specimen configuration and nominal dimensions (mm) for as-forged
surface finish specimens subjected to (a) cantilever bending fatigue, and
(b) rotating bending fatigue.













97



(a)


(b)


98

(c)

99

(d)

Figure 3-3: Flow charts showing (a) specimen preparation, and various testing
conditions for (b) specimens machined form rolled bar, (c) gas furnace
heated forged specimens, and (d) induction heated forged specimens.


100





Figure 3-4: Test setup used for reversed cantilever bending fatigue testing of as-
forged, shot-cleaned, and machined and polished specimens.


Apparatus 2
Apparatus 1
Specimen
Load Cell
Loading Direction
Universal Ball Joint
Roller Bearings

101
-1500
-1000
-500
0
500
1000
1500
-1500 -1000 -500 0 500 1000 1500
Measured Strain ()
C
a
l
c
u
l
a
t
e
d

S
t
r
a
i
n

(

)






.
Top Gage
Bottom Gage


Figure 3-5: Plot of calculated strain versus measured strain used to verify cantilever
bending test setup.










102




Figure 3-6: Four-point rotating bending fatigue testing machine.







103

(a)


(b)

Figure 3-7: Cross section of specimen gage section area prior to heat treatment
showing a mixture of ferrite and pearlite (Courtesy of Peter Bauerle of
Chrysler) for (a) induction heated specimens, and (b) gas furnace heated
specimens.
250 m
250 m

104

(a) (b)


(c) (d)

Figure 3-8: Magnification of gage section area showing decarburization for gas
furnace heated specimens with martensitic microstructure (Courtesy of
Peter Bauerle of Chrysler) at (a) 45 HRC, (b) 35 HRC, (c) 25 HRC, and
(d) 19 HRC.



Figure 3-9: Magnification of gage section area showing decarburization for induction
heated specimens at 45 HRC with martensitic microstructure (Courtesy of
Peter Bauerle of Chrysler).


105







Through the Trim Line

At 90 to the Trim Line


Figure 3-10: Grain flow resulting from forging process for induction heated specimens
(Courtesy of Peter Bauerle of Chrysler).




Figure 3-11: Magnified image showing surface irregularities for induction heated
specimens prior to heat treatment (Courtesy of Peter Bauerle of Chrysler).



106


Figure 3-12: Brinell hardness versus depth below the surface for induction heated
forged and gas furnace heated forged specimens.















B
r
i
n
e
l
l

H
a
r
d
n
e
s
s
,

H
B
Depth Below Surface (mm)
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
100
200
300
400
Legend
Gas 45 HRC
Gas 35 HRC
Gas 25 HRC
Gas 19 HRC
Induction 45 HRC
Induction 35 HRC
Induction 25 HRC
Induction 19 HRC

107
-30
-20
-10
0
10
20
30
0 2 4 6 8 10
(mm)
(

m
)

(a)
-30
-20
-10
0
10
20
30
0 2 4 6 8 10
(mm)
(

m
)

(b)

-30
-20
-10
0
10
20
30
0 2 4 6 8 10
(mm)
(

m
)

(c)

Figure 3-13: Roughness profiles for induction heated specimens (a) prior to heat
treatment, (b) as-forged at 45 HRC, and (c) as-forged and shot cleaned at
45 HRC.


108
-30
-20
-10
0
10
20
0 2 4 6 8 10
(mm)
(

m
)

(a)

-20
-10
0
10
20
30
0 2 4 6 8 10
(mm)
(

m
)

(b)

-30
-20
-10
0
10
20
0 2 4 6 8 10
(mm)
(

m
)

(c)

Figure 3-14: Roughness profiles for gas furnace heated specimens (a) prior to heat
treatment, (b) as-forged at 45 HRC, and (c) as-forged and shot cleaned at
45 HRC.






109

F
i
g
u
r
e

3
-
1
5
:




R
o
u
g
h
n
e
s
s

v
e
r
s
u
s

h
a
r
d
n
e
s
s

f
o
r

s
h
o
t

c
l
e
a
n
e
d

a
n
d

a
s
-
f
o
r
g
e
d

s
u
r
f
a
c
e
s

f
o
r

(
a
)

r
o
u
g
h
n
e
s
s

p
a
r
a
m
e
t
e
r

R
a
,

(
b
)




r
o
u
g
h
n
e
s
s

p
a
r
a
m
e
t
e
r

R
t
,

(
c
)

r
o
u
g
h
n
e
s
s

p
a
r
a
m
e
t
e
r

R
y
,

a
n
d

(
d
)

r
o
u
g
h
n
e
s
s

p
a
r
a
m
e
t
e
r

R
z
.



110
0
200
400
600
800
1000
1200
1400
1600
1800
0 1 2 3 4 5 6 7 8 9 10
Engineering Strain, (%)
E
n
g
i
n
e
e
r
i
n
g

S
t
r
e
s
s
,

(
M
P
a
)





















.
19 HRC
25 HRC
35 HRC
45 HRC

Figure 3-16: Monotonic stress-strain curves at different hardness levels for 10B40 steel.

600
700
800
900
1000
1100
1200
1300
1400
1500
1600
200 250 300 350 400 450
Brinell Hardness (HB)
T
e
n
s
i
l
e

S
t
r
e
n
g
t
h

(
M
P
a
)



.
Data
Estimation
S
u
= 3.45 (HB)


Figure 3-17: Tensile strength versus Brinell hardness for 10B40 steel.


111

(a) (b)


(c) (d)


(e) (f)


(h) (i)

Figure 3-18: Typical fracture surface of as-forged specimens subjected to fully-reversed
cantilever bending fatigue at (a) stress amplitude of 927 MPa at 45 HRC,
(b) stress amplitude of 301 MPa at 45 HRC, (c) stress amplitude of 610
MPa at 35 HRC, (d) stress amplitude of 251 MPa at 35 HRC, (e) stress
amplitude of 609 MPa at 25 HRC, (f) stress amplitude of 260 MPa at 25
HRC, (h) stress amplitude of 500 MPa at 19 HRC, and (i) stress amplitude
of 249 MPa at 19 HRC.



112














































F
i
g
u
r
e

3
-
1
9
:




D
i
s
p
l
a
c
e
m
e
n
t

a
m
p
l
i
t
u
d
e

v
e
r
s
u
s

n
o
r
m
a
l
i
z
e
d

c
y
c
l
e
s

i
n

c
a
n
t
i
l
e
v
e
r

b
e
n
d
i
n
g

f
a
t
i
g
u
e

t
e
s
t
s

f
o
r

4
5

H
R
C

s
p
e
c
i
m
e
n
s

(
a
)




a
s
-
f
o
r
g
e
d

w
i
t
h

i
n
d
u
c
t
i
o
n

h
e
a
t
,

(
b
)

a
s
-
f
o
r
g
e
d

w
i
t
h

g
a
s

f
u
r
n
a
c
e

h
e
a
t
,

(
c
)

a
s
-
f
o
r
g
e
d

w
i
t
h

i
n
d
u
c
t
i
o
n

h
e
a
t

a
n
d

s
h
o
t
-



c
l
e
a
n
e
d
,

a
n
d

(
d
)

a
s
-
f
o
r
g
e
d

w
i
t
h

g
a
s

f
u
r
n
a
c
e

h
e
a
t

a
n
d

s
h
o
t

c
l
e
a
n
e
d
.






113














































F
i
g
u
r
e

3
-
2
0
:




D
i
s
p
l
a
c
e
m
e
n
t

a
m
p
l
i
t
u
d
e

v
e
r
s
u
s

n
o
r
m
a
l
i
z
e
d

c
y
c
l
e
s

i
n

c
a
n
t
i
l
e
v
e
r

b
e
n
d
i
n
g

f
a
t
i
g
u
e

t
e
s
t
s

f
o
r

3
5

H
R
C

s
p
e
c
i
m
e
n
s

(
a
)


a
s
-
f
o
r
g
e
d

w
i
t
h

i
n
d
u
c
t
i
o
n

h
e
a
t
,

(
b
)

a
s
-
f
o
r
g
e
d

w
i
t
h

g
a
s

f
u
r
n
a
c
e

h
e
a
t
,

(
c
)

a
s
-
f
o
r
g
e
d

w
i
t
h

i
n
d
u
c
t
i
o
n

h
e
a
t

a
n
d

s
h
o
t
-

c
l
e
a
n
e
d
,

a
n
d

(
d
)

a
s
-
f
o
r
g
e
d

w
i
t
h

g
a
s

f
u
r
n
a
c
e

h
e
a
t

a
n
d

s
h
o
t

c
l
e
a
n
e
d
.







114














































F
i
g
u
r
e

3
-
2
1
:




D
i
s
p
l
a
c
e
m
e
n
t

a
m
p
l
i
t
u
d
e

v
e
r
s
u
s

n
o
r
m
a
l
i
z
e
d

c
y
c
l
e
s

i
n

c
a
n
t
i
l
e
v
e
r

b
e
n
d
i
n
g

f
a
t
i
g
u
e

t
e
s
t
s

f
o
r

2
5

H
R
C

s
p
e
c
i
m
e
n
s

(
a
)


a
s
-
f
o
r
g
e
d

w
i
t
h

i
n
d
u
c
t
i
o
n

h
e
a
t
,

(
b
)

a
s
-
f
o
r
g
e
d

w
i
t
h

g
a
s

f
u
r
n
a
c
e

h
e
a
t
,

(
c
)

a
s
-
f
o
r
g
e
d

w
i
t
h

i
n
d
u
c
t
i
o
n

h
e
a
t

a
n
d

s
h
o
t
-

c
l
e
a
n
e
d
,

a
n
d

(
d
)

a
s
-
f
o
r
g
e
d

w
i
t
h

g
a
s

f
u
r
n
a
c
e

h
e
a
t

a
n
d

s
h
o
t

c
l
e
a
n
e
d
.







115














































F
i
g
u
r
e

3
-
2
2
:




D
i
s
p
l
a
c
e
m
e
n
t

a
m
p
l
i
t
u
d
e

v
e
r
s
u
s

n
o
r
m
a
l
i
z
e
d

c
y
c
l
e
s

i
n

c
a
n
t
i
l
e
v
e
r

b
e
n
d
i
n
g

f
a
t
i
g
u
e

t
e
s
t
s

f
o
r

1
9

H
R
C

s
p
e
c
i
m
e
n
s

(
a
)


a
s
-
f
o
r
g
e
d

w
i
t
h

i
n
d
u
c
t
i
o
n

h
e
a
t
,

(
b
)

a
s
-
f
o
r
g
e
d

w
i
t
h

g
a
s

f
u
r
n
a
c
e

h
e
a
t
,

(
c
)

a
s
-
f
o
r
g
e
d

w
i
t
h

i
n
d
u
c
t
i
o
n

h
e
a
t

a
n
d

s
h
o
t
-

c
l
e
a
n
e
d
,

a
n
d

(
d
)

a
s
-
f
o
r
g
e
d

w
i
t
h

g
a
s

f
u
r
n
a
c
e

h
e
a
t

a
n
d

s
h
o
t

c
l
e
a
n
e
d
.





116








F
i
g
u
r
e

3
-
2
3
:




D
i
s
p
l
a
c
e
m
e
n
t

a
m
p
l
i
t
u
d
e

v
e
r
s
u
s

n
o
r
m
a
l
i
z
e
d

c
y
c
l
e
s

i
n

c
a
n
t
i
l
e
v
e
r

b
e
n
d
i
n
g

f
a
t
i
g
u
e

t
e
s
t
s

f
o
r

m
a
c
h
i
n
e
d

a
n
d

p
o
l
i
s
h
e
d


s
p
e
c
i
m
e
n
s

a
t

(
a
)

4
5

H
R
C
,

(
b
)

3
5

H
R
C
,

(
c
)

2
5

H
R
C
,

a
n
d

(
d
)

1
9

H
R
C
.




117
F
i
g
u
r
e

3
-
2
4
:




S
t
r
e
s
s
-
l
i
f
e

c
u
r
v
e
s

u
n
d
e
r

c
a
n
t
i
l
e
v
e
r

b
e
n
d
i
n
g

f
a
t
i
g
u
e

f
o
r

a
s
-
f
o
r
g
e
d

i
n
d
u
c
t
i
o
n

h
e
a
t
e
d

s
p
e
c
i
m
e
n
s

a
t

(
a
)

4
5

H
R
C
,

(
b
)


3
5

H
R
C
,

(
c
)

2
5

H
R
C
,

a
n
d

(
d
)

1
9

H
R
C
.



118

F
i
g
u
r
e

3
-
2
5
:




S
t
r
e
s
s
-
l
i
f
e

c
u
r
v
e
s

u
n
d
e
r

c
a
n
t
i
l
e
v
e
r

b
e
n
d
i
n
g

f
a
t
i
g
u
e

f
o
r

a
s
-
f
o
r
g
e
d

a
n
d

s
h
o
t

c
l
e
a
n
e
d

i
n
d
u
c
t
i
o
n

h
e
a
t
e
d

s
p
e
c
i
m
e
n
s

a
t


(
a
)

4
5

H
R
C
,

(
b
)

3
5

H
R
C
,

(
c
)

2
5

H
R
C
,

a
n
d

(
d
)

1
9

H
R
C
.



119

F
i
g
u
r
e

3
-
2
6
:




S
t
r
e
s
s
-
l
i
f
e

c
u
r
v
e
s

u
n
d
e
r

c
a
n
t
i
l
e
v
e
r

b
e
n
d
i
n
g

f
a
t
i
g
u
e

f
o
r

a
s
-
f
o
r
g
e
d

g
a
s

f
u
r
n
a
c
e

h
e
a
t
e
d

s
p
e
c
i
m
e
n
s

a
t

(
a
)

4
5

H
R
C
,


(
b
)

3
5

H
R
C
,

(
c
)

2
5

H
R
C
,

a
n
d

(
d
)

1
9

H
R
C
.



120


F
i
g
u
r
e

3
-
2
7
:




S
t
r
e
s
s
-
l
i
f
e

c
u
r
v
e
s

u
n
d
e
r

c
a
n
t
i
l
e
v
e
r

b
e
n
d
i
n
g

f
a
t
i
g
u
e

f
o
r

a
s
-
f
o
r
g
e
d

a
n
d

s
h
o
t

c
l
e
a
n
e
d

g
a
s

f
u
r
n
a
c
e

h
e
a
t
e
d

s
p
e
c
i
m
e
n
s




a
t

(
a
)

4
5

H
R
C
,

(
b
)

3
5

H
R
C
,

(
c
)

2
5

H
R
C
,

a
n
d

(
d
)

1
9

H
R
C
.



121

F
i
g
u
r
e

3
-
2
8
:




S
t
r
e
s
s
-
l
i
f
e

c
u
r
v
e
s

u
n
d
e
r

c
a
n
t
i
l
e
v
e
r

b
e
n
d
i
n
g

f
a
t
i
g
u
e

f
o
r

m
a
c
h
i
n
e
d

a
n
d

p
o
l
i
s
h
e
d

s
p
e
c
i
m
e
n
s

a
t

(
a
)

4
5

H
R
C
,

(
b
)

3
5




H
R
C
,

(
c
)

2
5

H
R
C
,

a
n
d

(
d
)

1
9

H
R
C
.



122
-1000
-800
-600
-400
-200
0
200
400
600
800
1000
54 55 56 57 58 59 60
Displacement (mm)
L
o
a
d

(
N
)
45 HRC
35 HRC
25 HRC
19 HRC

(a)
-1000
-800
-600
-400
-200
0
200
400
600
800
1000
54 55 56 57 58 59 60
Displacement (mm)
L
o
a
d

(
N
)
45 HRC
35 HRC
25 HRC
19 HRC

(b)

Figure 3-29: Midlife load versus displacement curves for highest load amplitude tests at
45 HRC, 35 HRC, 25 HRC, and 19 HRC for (a) induction heated forged
specimens, and (b) machined and polished specimens.




123



(a) (b)


(c) (d)

Figure 3-30: Typical fracture surface of as-forged specimens subjected to rotating
bending fatigue at (a) stress amplitude of 551 MPa at 35 HRC, (b) stress
amplitude of 229 MPa at 35 HRC, (c) stress amplitude of 502 MPa at 19
HRC, and (d) stress amplitude of 227 MPa at 19 HRC.















124




F
i
g
u
r
e

3
-
3
1
:




S
t
r
e
s
s
-
l
i
f
e

c
u
r
v
e
s

u
n
d
e
r

r
o
t
a
t
i
n
g

b
e
n
d
i
n
g

f
a
t
i
g
u
e

f
o
r

(
a
)

g
a
s

f
u
r
n
a
c
e

h
e
a
t
e
d

a
s
-
f
o
r
g
e
d

s
u
r
f
a
c
e

a
t

3
5

H
R
C
,

(
b
)




m
a
c
h
i
n
e
d

a
n
d

p
o
l
i
s
h
e
d

s
u
r
f
a
c
e

a
t

3
5

H
R
C
,

(
c
)

g
a
s

f
u
r
n
a
c
e

h
e
a
t
e
d

a
s
-
f
o
r
g
e
d

s
u
r
f
a
c
e

a
t

1
9

H
R
C
,

a
n
d

(
d
)




m
a
c
h
i
n
e
d

a
n
d

p
o
l
i
s
h
e
d

s
u
r
f
a
c
e

a
t

1
9

H
R
C
.



125






Chapter 4





Comparative Analysis and Discussion




This chapter discusses comparisons between the different types of surface finish
conditions including as-forged surface by two heating methods, machined and polished
surface, and as-forged and shot cleaned surfaces. The shot cleaned surface has an added
effect of compressive residual stress which is discussed. Also, rotating bending fatigue
test results are compared to reverse cantilever bending fatigue test results. In addition, the
effect of the forging flash or trim line is discussed. There is also a discussion of the grain
flow effect resulting from the forging process.

4.1 Cantilever Bending vs. Rotating Bending Fatigue
Figure 4-1 shows composite plots of both cantilever bending and rotating bending
fatigue data for as-forged as well as machined and polished surface conditions at 35 HRC
and 19 HRC. The results for the two loading conditions are very similar for both surface
conditions and at both hardness levels. In theory, smooth specimens under rotating
bending fatigue should experience shorter life compared to reverse cantilever bending
fatigue. This is due to the fact that the entire circumference of the gage section is
subjected to highest stress in rotating bending, as compared to two points of the gage

126
section in cantilever bending. As a result, there is a higher probability that an area
subjected to high stresses contains a defect under rotating bending conditions. However,
the as-forged surface condition results in defects in the entire circumference of the
specimen gage section. This may explain why there was less of an effect on the fatigue
strength at 10
6
cycles for the as-forged condition compared to the polished surface
condition between reverse cantilever and rotating bending fatigue.
The correction factor for size effect was given in Chapter 2 and repeated below
for convenience:
( )

s <
s s
=

mm 254 51 51 . 1
mm 51 79 . 2 762 . 0 /
157 . 0
107 . 0
d d
d d
k
b
(4.1)
d d
e
37 . 0 = (4.2)
Equation 4.1 is used to correct the endurance limit for rotating bending fatigue specimens
having different diameters. Equation 4.2 is used to calculate an equivalent diameter for
non-rotating bending and is then substituted into Eqn. 4.1. This results in less of a size
effect compared to rotating bending (about 10% increase in k
b
value for non-rotating
bending according to Eqn. 4.1 based on a diameter of 8 mm). It should be noted that the
difference between rotating and nonrotating bending has been reported to be typically
less that 5% for commercial materials [35].

4.2 Forging Flash Effect
In some of the cantilever bending tests the effect of the flash line was investigated
by positioning the flash line in several gas furnace heated forged specimens at the
maximum stress location. Gas furnace heated specimens were chosen for this comparison

127
instead of induction heated specimens because the flash was uniformly trimmed for gas
furnace heated specimens, as would be done in industry (see Figure 3-1 showing the flash
line for both heating conditions). The effect was found to be insignificant (see Figure 4-
2). Similarly, in rotating bending fatigue tests of the as-forged surface finish specimens
with gas furnace heating, cracks did not initiate at the flash line in about half of the tests.
In addition, in these rotating bending tests the crack initiation location was independent
of the fatigue life, as the specimens with cracks initiating at the flash did not always fail
sooner than the specimens with cracks not initiating at the flash line. Therefore, as
discussed earlier, it can be assumed that the forging flash had little or no effect on fatigue
behavior of the as-forged surface finish specimens. The reason the flash had little effect
on the fatigue behavior may be attributed to the additional material from the flash
reducing the stress, as well as due to the surface roughness and discontinuities which may
outweigh any stress concentration effects of the flash.

4.3 Effect of Hardness
Figure 4-3 shows how hardness affects the fatigue behavior of the machined and
polished forged specimens. The curves are in a descending order starting with the highest
hardness of 45 HRC. As the hardness decreases, the fatigue strength also decreases, as
expected. Fatigue strength of steels at 10
6
cycles for smooth polished specimens is often
estimated as half of the ultimate tensile strength, up to a tensile strength of 1400 MPa. In
the current study, cantilever bending fatigue strength at 10
6
cycles to tensile strength
ratios for smooth polished specimens were 0.52, 0.57, 0.59, and 0.64 for 45 HRC, 35
HRC, 25 HRC, and 19 HRC hardness levels, respectively. However, there was little to no

128
increase in fatigue strength at 10
6
cycles as tensile strength increases for both as-forged
and shot cleaned surface conditions of the forged specimens using either gas furnace
heating or induction heating, as seen in Figures 4-4 and 4-5.
Figure 4-4 shows the S-N curves for the as-forged cantilever bending specimens
with gas furnace and induction heating. Although the S-N curves are generally in
descending order with decreased hardness, fatigue limit values are not in descending
order with hardness. At 10
6
cycles the fatigue strength for forged specimens subjected to
gas furnace heating at 45 HRC, 35 HRC, 25 HRC, and 19 HRC are similar. In contrast, at
10
6
cycles the fatigue strength for 35 HRC, 25 HRC, and 19 HRC specimens subjected to
induction heating are in descending order; however the fatigue strength of the 45 HRC
specimens is lower than that of the 19 HRC specimens. This implies the surface condition
is the governing factor controlling fatigue behavior at long life, rather than hardness.
Figure 4-5 shows the effect of hardness for the shot cleaned specimens with gas furnace
and induction heat. The differences observed between shot-cleaned specimens of
different hardness are very similar to specimens having as-forged surface condition
without shot cleaning.
There are at least two factors contributing to this behavior of the as-forged surface
finish specimens. First, surface decarburization in the as-forged specimens was
significant. This resulted in a reduction of hardness (i.e. strength) in the decarburized
area. It can be seen from Figure 3-12 in a plot of hardness versus depth below the surface
that the gas furnace heated specimens had a greater depth of decarburization than the
induction heated specimens. The higher the hardness, the more significant the difference
between the surface and the core hardness values. Secondly, the rough surface condition

129
tends to have more of a detrimental effect on a harder material, as harder materials are
more sensitive to stress concentrations. The forged specimens had surface discontinuities
(see Figure 3-11). As a result, most of the life of the as-forged specimens consisted of
microcrack growth. It should be noted that the gas furnace heated specimens had deeper
surface discontinuities than the induction heated specimens. The greater depth of
decarburization and surface discontinuities for gas furnace heated specimens can explain
why the fatigue strength at 10
6
cycles for the four hardness levels tested were closer to
each other than the induction heated specimens.

4.4 Effect of Forging Heating Method
As previously mentioned forging with induction heating resulted in significantly
less decarburization as compared to gas furnace heating (see Figures 3-7 through 3-9). In
addition, specimens subjected to gas furnace heating contained surface cracks or
discontinuities almost twice the depth of the induction heated specimens (see Figure 3-11
for an example of surface discontinuities in induction heated specimens). Although the
induction heated specimens had higher measured surface roughness (see Figure 3-15),
this is believed to be due to the shot cleaning process used to clean them prior to heat
treatment. The gas furnace heated specimens were subjected to a less intense cleaning
than induction heated forged specimens.
The differences in surface condition of the forged specimens heated by the two
different methods resulted in slightly different fatigue behaviors, as seen in Figures 4-6
and 4-7 for the as-forged and the shot cleaned conditions, respectively. Although the
induction heated specimens had increased surface roughness, at 35 HRC, 25 HRC, and 19

130
HRC they still exhibited improved fatigue behavior compared to the gas furnace heated
specimens. As a result, the main cause of difference in fatigue behavior between the two
heating methods is believed to be due to decarburization and surface discontinuities (i.e.
microcracks), not surface roughness. It should be noted that the effect of heating method
was most significant at the 35 HRC level for both the as-forged and shot cleaned
specimens.
At 45 HRC, the gas furnace heated forged specimens had higher fatigue strength
than induction heated forged specimens at 10
6
cycles. This could be due to the fact that
the surface condition has more of an effect on a harder material. The gas furnace heating
resulted in significant decarburization causing about 50% reduction of hardness at the
surface for 45 HRC specimens. The induction heated specimens had only around 15%
reduction in hardness at the surface where the surface defects were located.

4.5 Effect of Shot Cleaning
As forged parts are commonly shot-cleaned after forging, the effect of shot
cleaning on fatigue behavior was also evaluated by shot cleaning some specimens after
heat treatment. The shot cleaning process used was representative of typical processes
used in forging industry. Shot cleaning has benefit of compressive residual stresses at the
surface, but also causes increased surface roughness. In the LCF region, the residual
stresses can relax and cause shorter fatigue life due to increased surface roughness.
However, in the HCF region, the effect of compressive residual surface stresses
outweighs the effect of increased surface roughness and results in longer fatigue lives.
Figures 4-8 and 4-9 show the effect of shot cleaning on as-forged specimens with

131
induction and gas furnace heat forging, respectively. Test results show shot cleaning
improved fatigue behavior in the HCF region due to the surface compressive residual
stress it imparts on the part. Shot cleaning resulted in 8% to 17% increase in fatigue
strength at 10
6
cycles, depending on hardness level. The increase in fatigue strength was
similar between gas furnace heated and induction heated specimens. There was little to
no effect of shot cleaning at short lives of around 10
4
cycles.

4.6 Forging Grain Flow Effect
The forging process resulted in the typical grain flow, as seen in Figure 3-12. The
grain flow also reorients inclusions along the grain flow lines, which helps to improve
fatigue behavior. The effect of grain flow was evaluated by machining specimens from
the same rolled bars used for the forging process and then comparing the test results to
those obtained from forged specimens with the forged surface machined away. Figure 4-
10 shows the grain flow effect for 45 HRC and 25 HRC hardness specimens. It can be
seen that there is little to no effect on the type of steel and specimen used in this
investigation. There are two possible reasons for this. First, the rolled bar itself has some
grain flow from the rolling process. Second, the specimen geometry with smooth gage
section geometry cannot exhibit possible beneficial effect of grain flow of a forged part,
compared to a machined part where the machining process can cut across the grains (for
example, a forged stepped shaft versus a machined stepped shaft).




132
4.7 Effect of Surface Condition
Figure 2-26 shows schematic representation of S-N curves for smooth and
polished surface and for forged surface. The surface finish factor for a given hardness is
calculated by dividing the fatigue strength of the as-forged surface finish part by the
fatigue strength of the smooth and polished surface part at long life (typically at 10
6

cycles to failure for steels). Table 4.1 summarizes the surface finish factors for
experimental and historical data at four hardness levels (45 HRC, 35 HRC, 25 HRC, and
19 HRC) for fatigue lives ranging from 10
4
to 10
6
cycles. The surface finish factors in
Table 4.1 for historical data obtained from [2] for life less than 10
6
cycles were calculated
using S-N lines for polished specimens and as-forged specimens that intersect at 90% of
the ultimate tensile strength at 10
3
cycles and extend to a fatigue limit at 10
6
cycles, as
recommended in [13]. The surface finish factors for experimental data were obtained by
dividing the fatigue strength given by the S-N line for the as-forged surface specimens by
the fatigue strength given by S-N line for the machined and polished surface specimens.
Figure 4-11 compares cantilever bending fatigue behaviors of as-forged and shot
cleaned surface conditions to machined and polished surface condition at the 45 HRC
hardness level. It can be seen in this figure that there is a significant difference in fatigue
behavior between the as-forged (see Figure 4-11(a)) or the as-forged and shot cleaned
(see Figure 4-11(b)) surfaces, compared to the machined and polished surface. One main
cause of the difference is the surface roughness. However, the effect of surface roughness
is thought to be limited to the crack initiation stage [48]. In addition to surface roughness,
the forging process resulted in discontinuities as deep as 0.06 mm for induction heated
forged specimens and 0.1 mm for the gas furnace heated forged specimens. As a result of

133
the surface defects, fatigue life for specimens with the as-forged surface condition
consists of mostly short crack growth. Surface decarburization is another cause of the
difference observed between the as-forged surface and machined and polished surface
conditions, as it lowered the strength of the surface.
Figure 4-11 also includes a dotted line which represents the prediction of the
fatigue behavior of the as-forged specimen based on the old surface finish effect data.
The sloped portion of this line was obtained by multiplying the stress value of the S-N
line for the machined and polished specimens at 10
6
cycles by a surface finish factor of
0.23 for 45 HRC (421 HB) from [2], and connecting it to the 90% of the ultimate tensile
strength at 10
3
cycles. The surface finish factor calculated at 10
6
cycles for the 45 HRC
specimens from data collected in the current investigation is about 0.32 for gas furnace
heated forged specimens and 0.27 for induction heated forged specimens (see Table 4.1
for calculated surface finish factors). This results in a fatigue strength at 10
6
cycles that is
about 15% to 30% greater than the prediction based on historical surface finish data and
about three times longer fatigue life. The surface finish factors for the shot cleaned gas
furnace heated and induction heated forged specimens are 0.39 and 0.34, respectively,
resulting in 30% to 40% higher fatigue strength than predictions based on old data and
longer life by almost an order of magnitude.
Figure 4-12 shows the effect of surface condition for the 35 HRC hardness
specimens. Similar to the 45 HRC hardness specimens, surface roughness and
decarburization are the main reasons for the differences between the fatigue behaviors of
the as-forged and the machined and polished specimens. This figure also includes the
prediction line for the as-forged surface condition based on the surface finish factor from

134
[2] (a factor of 0.30 for 35 HRC or 327 HB). However, the surface finish factor
calculated at 10
6
cycles for the 35 HRC specimens from data in the current investigation
is about 0.42 and 0.47 for gas furnace heat and induction heat forged specimens,
respectively, resulting in 28% to 36% higher fatigue strength at 10
6
cycles and about
three to five times longer fatigue life. The surface finish factors for the shot cleaned gas
furnace heat and induction heat forged specimens are 0.45 and 0.56 (see Table 4.1),
respectively. This results in predicted fatigue strengths at 10
6
cycles 35% to 45% greater
compared to fatigue strengths predicted using the historical data and longer life by around
an order of magnitude.
Figures 4-13 and 4-14 show cantilever bending fatigue data and S-N curves for
the as-forged (Figures 4-13(a) and 4-14(a)) or the as-forged and shot cleaned (Figures 4-
13(b) and 4-14(b)) surfaces, compared to the machined and polished surface conditions
of the 25 HRC and the 19 HRC hardness specimens, respectively. Surface roughness and
decarburization are again the main reasons for the differences observed between the as-
forged or as-forged and shot cleaned and the machined and polished specimen fatigue
data and S-N curves. The S-N prediction lines for the fatigue behavior of as-forged
specimens based on data from [2] shown in these figures indicate overly conservative
predictions. The surface finish factor at 10
6
cycles in [2] for 25 HRC (253 HB) and 19
HRC (220 HB) is 0.35 and 0.38, respectively. This factor is about 0.51 and 0.56 for gas
furnace heated forged and induction heated forged 25 HRC as-forged surface specimens,
respectively (see Table 4.1), and 0.58 and 0.61 for the as-forged and shot cleaned surface
for gas furnace and induction heated forging. The surface finish factors for the 19 HRC
as-forged surface specimens are 0.55 and 0.59 for gas furnace heated and induction

135
heated forging, respectively. The surface finish factors for the as-forged and shot cleaned
gas furnace heated and induction heated specimens are 0.63 and 0.69, respectively. This
results in a fatigue strength at 10
6
cycles that is about 33% greater than the predicted
value (using historical data) at either hardness level for as-forged gas furnace and
induction heated forgings, and around five to ten times longer fatigue life. The surface
finish factors for the as-forged and shot cleaned specimens results in fatigue strength (10
6

cycles), which is about 45% greater than prediction using historical data for induction
heated specimens at both 25 HRC and 19 HRC hardness levels and about 40% greater
than predictions for gas furnace heated specimens at 25 HRC and 19 HRC hardness
levels.
Comparisons of Figures 4-11 through 4-14 indicate the higher hardness materials
exhibit more of a difference in fatigue strength at 10
6
cycles between the as-forged and
machined and polished surface conditions. This can also be observed from [2], where the
surface finish factor reduces (i.e. more fatigue strength reduction) with increasing
hardness. This is explained by higher sensitivity of higher hardness materials to stress
concentrations, such as those induced by surface roughness and decarburized layer
defects in the as-forged surface specimens. It should be noted that there is little to no
effect of surface finish effect at 10
3
cycles, due to inelastic behavior.
Figure 4-15 shows the surface finish factor at 10
6
cycles versus tensile strength or
Brinell hardness. This figure includes a curved line based on data taken from [2], as well
as surface finish factor values calculated using fatigue data from the current investigation.
There are four sets of points in this figure representing as-forged gas furnace heated
specimens, as-forged and shot cleaned gas furnace heated specimens, as-forged induction

136
heated specimens, and as-forged and shot cleaned induction heated specimens. In
addition, there are dashed lines fit to the as-forged gas furnace and induction heated
specimens and the as-forged and shot cleaned gas furnace and induction heated
specimens. It can be seen that there is a significant difference between data from the
current investigation and the commonly used historical data taken from [2]. The
comparison indicates the historical data are overly conservative.
Figure 4-16 shows a plot of predicted fatigue life using historical data versus
experimental fatigue life, which summarizes the fatigue test data for the as-forged, and
as-forged and shot cleaned specimens of both forging heating methods used and at all
four hardness levels. This figure includes a 45 degree line representing perfect agreement,
a factor of three difference line, a factor of 10 difference line, and a factor of 100
difference line. It can be seen in this figure that the predicted fatigue lives using historical
data are significant lower than experimental fatigue lives and, therefore, very
conservative.













137
Table 4.1: Summary of calculated surface finish factors for as-forged gas furnace
heated, as-forged induction heated, as-forged and shot cleaned gas furnace
heated, and as-forged and shot cleaned induction heated specimens, as
well as historical surface finish factors.












220 HB
(19 HRC)
253 HB
(25 HRC)
327 HB
(35 HRC)
421 HB
(45 HRC)
1.E+04 0.95 1.09 0.84 0.83
3.E+04 0.85 0.91 0.74 0.71
1.E+05 0.76 0.75 0.65 0.60
3.E+05 0.68 0.63 0.52 0.43
1.E+06 0.55 0.51 0.42 0.34
1.E+04 0.913 0.845
3.E+04 0.840 0.723
1.E+05 0.766 0.610
3.E+05 0.704 0.522
1.E+06 0.606 0.430
1.E+04 0.93 1.00 0.83 0.82
3.E+04 0.87 0.88 0.77 0.75
1.E+05 0.80 0.76 0.70 0.68
3.E+05 0.74 0.67 0.59 0.52
1.E+06 0.63 0.58 0.45 0.39
1.E+04 0.95 1.02 0.85 0.84
3.E+04 0.86 0.89 0.79 0.70
1.E+05 0.78 0.76 0.73 0.57
3.E+05 0.71 0.66 0.61 0.39
1.E+06 0.59 0.56 0.46 0.26
1.E+04 0.95 1.01 0.83 0.81
3.E+04 0.90 0.89 0.81 0.72
1.E+05 0.85 0.79 0.79 0.63
3.E+05 0.80 0.70 0.70 0.47
1.E+06 0.69 0.61 0.56 0.34
1.E+04 0.72 0.70 0.67 0.61
3.E+04 0.62 0.59 0.55 0.49
1.E+05 0.52 0.49 0.45 0.38
3.E+05 0.44 0.42 0.37 0.30
1.E+06 0.38 0.35 0.30 0.23
GF Rotating
Condition N
f
Surface Finish Factor (k
s
)
Gas Furnace
Heating As-
Forged
Historical Data
Gas Furnace
Heating As-
Forged and
Shot Cleaned
Induction
Heating As-
Forged
Induction
Heating As-
Forged and
Shot Cleaned

138


(a)

(b)

Figure 4-1: Cantilever bending versus rotating bending for as-forged and polished
surface conditions at (a) 35 HRC, and (b) 19 HRC.
S
t
r
e
s
s

A
m
p
l
i
t
u
d
e
,

a
(
M
P
a
)
Cycles to Failure, N
f
10
3
10
4
10
5
10
6
10
7
100
200
300
400
500
600
700
800
900
1000
2000
(2)
(2)
(3)
(3)
Legend
GF19
GF19 (Rotating)
M19
M19 (Rotating)
S
t
r
e
s
s

A
m
p
l
i
t
u
d
e
,

a
(
M
P
a
)
Cycles to Failure, N
f
10
3
10
4
10
5
10
6
10
7
100
200
300
400
500
600
700
800
900
1000
2000
Legend
GF35
GF35 (Rotating)
M35
M35 (Rotating)

139

(a)

(b)

Figure 4-2: Forging flash effect on fatigue behavior of as-forged specimens with gas
furnace heating in cantilever bending tests for (a) 45 HRC, and (b) 25
HRC.
S
t
r
e
s
s

A
m
p
l
i
t
u
d
e
,

a
(
M
P
a
)
Cycles to Failure, N
f
10
3
10
4
10
5
10
6
10
7
100
200
300
400
500
600
700
800
900
1000
2000
(2)
(2)
(4)
Legend
GF25
GF25 (Flash)
S
t
r
e
s
s

A
m
p
l
i
t
u
d
e
,

a
(
M
P
a
)
Cycles to Failure, N
f
10
3
10
4
10
5
10
6
10
7
100
200
300
400
500
600
700
800
900
1000
2000
(2)
Legend
GF45
GF45 (Flash)

140



Figure 4-3: Effect of hardness on fatigue behavior of machined and polished
specimens in cantilever bending.




S
t
r
e
s
s

A
m
p
l
i
t
u
d
e
,

a
(
M
P
a
)
Cycles to Failure, N
f
10
3
10
4
10
5
10
6
10
7
100
200
300
400
500
600
700
800
900
1000
2000
Legend
M45
M35
M25
M19

141

(a)

(b)

Figure 4-4: Effect of hardness on cantilever bending fatigue behavior for as-forged
specimens with (a) induction heating, and (b) gas furnace heating.
S
t
r
e
s
s

A
m
p
l
i
t
u
d
e
,

a
(
M
P
a
)
Cycles to Failure, N
f
10
3
10
4
10
5
10
6
10
7
100
200
300
400
500
600
700
800
900
1000
2000
Legend
IF45
IF35
IF25
IF19
S
t
r
e
s
s

A
m
p
l
i
t
u
d
e
,

a
(
M
P
a
)
Cycles to Failure, N
f
10
3
10
4
10
5
10
6
10
7
100
200
300
400
500
600
700
800
900
1000
2000
Legend
GF45
GF35
GF25
GF19

142

(a)

(b)

Figure 4-5: Effect of hardness on cantilever bending fatigue behavior for shot cleaned
forged specimens with (a) induction heating, and (b) gas furnace heating.
10
3
10
4
10
5
10
6
10
7
100
200
300
400
500
600
700
800
900
1000
2000
S
t
r
e
s
s

A
m
p
l
i
t
u
d
e
,

a
(
M
P
a
)
Cycles to Failure, N
f
Legend
IS45
IS35
IS25
IS19
10
3
10
4
10
5
10
6
10
7
100
200
300
400
500
600
700
800
900
1000
2000
S
t
r
e
s
s

A
m
p
l
i
t
u
d
e
,

a
(
M
P
a
)
Cycles to Failure, N
f
Legend
GS45
GS35
GS25
GS19

143

F
i
g

4
-
6
:





E
f
f
e
c
t

o
f

h
e
a
t
i
n
g

m
e
t
h
o
d

o
n

c
a
n
t
i
l
e
v
e
r

b
e
n
d
i
n
g

f
a
t
i
g
u
e

f
o
r

a
s
-
f
o
r
g
e
d

s
u
r
f
a
c
e

s
p
e
c
i
m
e
n
s

a
t

(
a
)

4
5

H
R
C
,

(
b
)

3
5




H
R
C
,

(
c
)

2
5

H
R
C
,

a
n
d

(
d
)

1
9

H
R
C
.


144

F
i
g

4
-
7
:





E
f
f
e
c
t

o
f

h
e
a
t
i
n
g

m
e
t
h
o
d

o
n

c
a
n
t
i
l
e
v
e
r

b
e
n
d
i
n
g

f
a
t
i
g
u
e

f
o
r

s
h
o
t

c
l
e
a
n
e
d

f
o
r
g
e
d

s
u
r
f
a
c
e

s
p
e
c
i
m
e
n
s

a
t

(
a
)

4
5

H
R
C
,




(
b
)

3
5

H
R
C
,

(
c
)

2
5

H
R
C
,

a
n
d

(
d
)

1
9

H
R
C
.



145

F
i
g

4
-
8
:





E
f
f
e
c
t

o
f

s
h
o
t

c
l
e
a
n
i
n
g

o
n

c
a
n
t
i
l
e
v
e
r

b
e
n
d
i
n
g

f
a
t
i
g
u
e

f
o
r

i
n
d
u
c
t
i
o
n

h
e
a
t
e
d

f
o
r
g
e
d

s
p
e
c
i
m
e
n
s

a
t

(
a
)

4
5

H
R
C
,

(
b
)


3
5

H
R
C
,

(
c
)

2
5

H
R
C
,

a
n
d

(
d
)

1
9

H
R
C
.



146


F
i
g

4
-
9
:

E
f
f
e
c
t

o
f

s
h
o
t

c
l
e
a
n
i
n
g

o
n

c
a
n
t
i
l
e
v
e
r

b
e
n
d
i
n
g

f
a
t
i
g
u
e

f
o
r

g
a
s

f
u
r
n
a
c
e

h
e
a
t
e
d

f
o
r
g
e
d

s
p
e
c
i
m
e
n
s

a
t

(
a
)

4
5

H
R
C
,

(
b
)




3
5

H
R
C
,

(
c
)

2
5

H
R
C
,

a
n
d

(
d
)

1
9

H
R
C
.



147


Figure 4-10: Comparison of cantilever bending fatigue behavior between forged
specimens with forged surface removed (circular symbols) and machined
specimens from rolled bars (triangular symbols) at 45 HRC and 25 HRC
hardness levels.








S
t
r
e
s
s

A
m
p
l
i
t
u
d
e
,

a
(
M
P
a
)
Cycles to Failure, N
f
10
3
10
4
10
5
10
6
10
7
100
200
300
400
500
600
700
800
900
1000
2000
(2)
(2)
(2)
(4)
Legend
M45
R45
M25
R25

148

(a)

(b)

Figure 4-11: Surface finish effect on cantilever bending fatigue behavior at 45 HRC
hardness level with induction and gas furnace heating for (a) as-forged,
and (b) as-forged and shot cleaned conditions.

10
3
10
4
10
5
10
6
10
7
100
200
300
400
500
600
700
800
900
1000
2000
S
t
r
e
s
s

A
m
p
l
i
t
u
d
e
,

a
(
M
P
a
)
Cycles to Failure, N
f
Legend
GS45
IS45
M45
Prediction Based on Old Data
10
3
10
4
10
5
10
6
10
7
100
200
300
400
500
600
700
800
900
1000
2000
S
t
r
e
s
s

A
m
p
l
i
t
u
d
e
,

a
(
M
P
a
)
Cycles to Failure, N
f
Legend
GF45
IF45
M45
Prediction Based on Old Data

149

(a)

(b)

Figure 4-12: Surface finish effect on cantilever bending fatigue behavior at 35 HRC
hardness level with induction and gas furnace heating for (a) as-forged,
and (b) as-forged and shot cleaned conditions.


10
3
10
4
10
5
10
6
10
7
100
200
300
400
500
600
700
800
900
1000
2000
S
t
r
e
s
s

A
m
p
l
i
t
u
d
e
,

a
(
M
P
a
)
Cycles to Failure, N
f
Legend
GS35
IS35
M35
Prediction Based on Old Data
10
3
10
4
10
5
10
6
10
7
100
200
300
400
500
600
700
800
900
1000
2000
S
t
r
e
s
s

A
m
p
l
i
t
u
d
e
,

a
(
M
P
a
)
Cycles to Failure, N
f
Legend
GF35
IF35
M35
Prediction Based on Old Data

150

(a)

(b)

Figure 4-13: Surface finish effect on cantilever bending fatigue behavior at 25 HRC
hardness level with induction and gas furnace heating for (a) as-forged,
and (b) as-forged and shot cleaned conditions.


10
3
10
4
10
5
10
6
10
7
100
200
300
400
500
600
700
800
900
1000
2000
S
t
r
e
s
s

A
m
p
l
i
t
u
d
e
,

a
(
M
P
a
)
Cycles to Failure, N
f
Legend
GS25
IS25
M25
Prediction Based on Old Data
10
3
10
4
10
5
10
6
10
7
100
200
300
400
500
600
700
800
900
1000
2000
S
t
r
e
s
s

A
m
p
l
i
t
u
d
e
,

a
(
M
P
a
)
Cycles to Failure, N
f
Legend
GF25
IF25
M25
Prediction Based on Old Data

151

(a)

(b)

Figure 4-14: Surface finish effect on cantilever bending fatigue behavior at 19 HRC
hardness level with induction and gas furnace heating for (a) as-forged,
and (b) as-forged and shot cleaned conditions.


10
3
10
4
10
5
10
6
10
7
100
200
300
400
500
600
700
800
900
1000
2000
S
t
r
e
s
s

A
m
p
l
i
t
u
d
e
,

a
(
M
P
a
)
Cycles to Failure, N
f
Legend
GS19
IS19
M19
Prediction Based on Old Data
10
3
10
4
10
5
10
6
10
7
100
200
300
400
500
600
700
800
900
1000
2000
S
t
r
e
s
s

A
m
p
l
i
t
u
d
e
,

a
(
M
P
a
)
Cycles to Failure, N
f
Legend
GF19
IF19
M19
Prediction Based on Old Data

152

(a)

(b)

Figure 4-15: Composite plot of surface finish factor at 10
6
cycles versus tensile strength
or hardness showing differences between induction and gas furnace
heating forging and old and current data with (a) as-forged surface, and (b)
as-forged and shot cleaned surface.


600 800 1000 1200 1400 1600
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
120 160 200 240 280 320 360 400 440 480 520
Brinell Hardness (HB)
Tensile Strength (MPa)
S
u
r
f
a
c
e

F
i
n
i
s
h

F
a
c
t
o
r
,

k
s
Legend
Shot Cleaned Induction Heating
Shot Cleaned Gas Furnace Heating
Historical Data
600 800 1000 1200 1400 1600
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
120 160 200 240 280 320 360 400 440 480 520
Brinell Hardness (HB)
Tensile Strength (MPa)
S
u
r
f
a
c
e

F
i
n
i
s
h

F
a
c
t
o
r
,

k
s
Legend
As-Forged Induction Heating
As-Forged Gas Furnace Heating
Historical Data

153
Experimental Fatigue Life (Cycles)
10
3
10
4
10
5
10
6
10
7
P
r
e
d
i
c
t
e
d

F
a
t
i
g
u
e

L
i
f
e

B
a
s
e
d

o
n

H
i
s
t
o
r
i
c
a
l

D
a
t
a

(
C
y
c
l
e
s
)
10
3
10
4
10
5
10
6
10
7


Figure 4-16: Predicted fatigue life based on historical data (cycles) versus experimental
fatigue life (cycles) for as-forged, and as-forged and shot cleaned
specimens at 45 HRC, 35 HRC, 25 HRC, and 19 HRC hardness levels.













19 HRC 25 HRC 35 HRC 45 HRC
Induction Heat
Induction Heat and Shot Cleaned
Gas Furnace Heat
Gas Furnace Heat and Shot Cleaned

154




Chapter 5





Fracture Mechanics Analysis




The as-forged surface is accompanied by surface roughness as well as surface
defects or discontinuities (see Figure 3-11). Some of the specimens in the current study
had discontinuities or cracks as deep as 0.1 mm. As a result, fatigue life of the as-forged
specimen may be considered dominated by short crack growth. The stress-life approach
used to quantify fatigue behavior for this investigation does not differentiate between
crack initiation and short crack growth life. Fracture mechanics analysis, on the other
hand, only considers crack growth life and can help to further explain the fatigue
behavior observed. This chapter discusses small crack growth behavior, an estimation of
crack growth rate, and prediction of fatigue life for as-forged bending specimens using a
fracture mechanics approach.

5.1 Crack Growth Rate Estimation
5.1.1 Crack growth behavior and rate
During a few cantilever bending fatigue tests, the length of the crack was
monitored. As the crack grew, the length was measured and the cycle number and
displacement amplitude were recorded. It should be noted that displacement amplitude

155
was recorded in all tests at regular intervals by the computer software controlling the test.
As the crack grew the displacement amplitude changed due to reduced stiffness (see
Figures 3-19 through 3-23). This change in displacement amplitude was then related to
the crack length at a given cycle (see Figure 5-1a). Similar behavior was observed
between data for the 45 HRC hardness and 35 HRC hardness specimens, and between the
25 HRC hardness and 19 HRC hardness specimens. As a result, one line was fit to the
four sets of data for the 45 HRC hardness and 35 HRC hardness specimens and another
line was fit to the four sets of data for the 25 HRC hardness and 19 HRC hardness
specimens. The difference between the two lines can be attributed to increased plasticity
at the crack tip (in the LCF region) for the lower hardness specimens. Using this
information, it was possible to calculate what change in displacement amplitude
corresponded to a certain length of crack.
The bending specimens experienced semi-elliptical cracks (aspect ratio of about
0.48) starting from the location of maximum stress, as seen in Figure 3-18. The
specimens tested at higher stress amplitudes show shorter crack depths compared to
specimens tested at low stress amplitudes. In addition, there are usually two cracks with
one at the top and one at the bottom for specimens tested at high stress amplitudes. The
specimens tested at low stress amplitudes had deeper cracks and often only had a crack
grow from one side (either top or bottom).
Figure 5-2 shows a plot of crack depth versus crack length, measured after
fracture, for several specimens at each of the four hardness levels. This figure includes a
best fit line which gives a relationship between crack depth and crack length defined as:
b a 48 . 0 = (5.1)

156
where, a is the crack depth and b is the crack length. Figure 5-3(a) shows how the crack
depth, a, and length, b, are defined on the fractured surface. Next, the relationship
between crack depth and crack length was used to estimate the crack depth for each cycle
that the length was recorded. Then it was possible to relate crack depth to a change in
displacement amplitude during fatigue tests, as seen in Figure 5-1(b). The crack length
and crack depth versus change in displacement amplitude having relationships given by:
368 . 0
) ( 98 . 5
a
b o A = For 45 HRC and 35 HRC (5.2)

368 . 0
) ( 87 . 2
a
a o A = For 45 HRC and 35 HRC (5.3)
324 . 0
) ( 58 . 6
a
b o A = For 25 HRC and 19 HRC (5.4)
324 . 0
) ( 16 . 3
a
a o A = For 25 HRC and 19 HRC (5.5)
where
a
o A is the change in displacement amplitude (with initial displacement amplitude
as the reference). It should be noted that Eqns. 5.2 and 5.3, as well as 5.4 and 5.5 are
related by Eqn. 5.1. Equations 5.2 and 5.3 are based on a best fit line to data from four
tests, two tests at each hardness level (45 HRC and 35 HRC), with stress amplitudes
ranging from 900 to 800 MPa. Equations 5.4 and 5.5 are also based on a best fit line to
data from four tests, two tests at each hardness level (25 HRC and 19 HRC), with stress
amplitudes ranging from 550 to 500 MPa. These ranges of stress are in the LCF range for
all four hardness levels. Figure 5-4 shows plots of change in displacement amplitude
versus normalized cycles.
As the change in displacement amplitude was measured periodically, the crack
depth could be calculated for each recorded cycle, which allowed the calculation of crack
growth rate (da/dN). Figure 5-5 shows plots of calculated crack depth, using Eqns. 5.3
and 5.5, versus normalized cycles representing the basis on which (da/dN) was

157
calculated. It should be noted that Eqns. 5.3 and 5.5 were only used to calculate crack
depths up to 3 mm. It can be seen in Figure 5-5 that the crack depth is increasing rapidly
after 3 mm (i.e. unstable crack growth).
The stress intensity factor range (K) was calculated for a given cycle using the
following equation:
( )
w
a
f a K
2
1
) (t o A = A (5.6)
where o A is the applied bending stress amplitude and ( )
w
a
f is the geometry factor. For
fully-reversed bending,
a
o o o = = A
max
, as the compression part of the cycle is not taken
into account. Geometry factors are unique to the geometry of the specimen and the crack,
as well as type of loading applied. The geometry factor used in the following analysis
was a constant equal to 0.728 which is for a semi-circular crack in a round bar [33]. The
round surface of the specimens can make a semi-circular crack appear to be more of a
semi-elliptical shape (see Figure 5-3(b)). It should be noted that the geometry factor used
was only applicable under the following condition:
35 . 0 <
d
a
(5.7)
where d is the diameter of the specimen, therefore, up to a crack depth of about 3 mm.

5.1.2 Crack growth rate versus stress intensity factor range
Figure 5-6 shows a schematic plot of crack growth rate (da/dN) versus stress
intensity factor range (K) in a log-log plot. It can be seen that the curve has a sigmoidal
shape with three regions. Region I, known as the near threshold region, typically has
crack growth rates below 10
-8
m/cycle. The threshold value (K
th
), the value of stress

158
intensity factor range at which cracks do not grow, is also shown in this region. Region
II, also known as the Paris equation region, is the linear portion of the curve represented
mathematically by:

n
K A
dN
da
) (A =
(5.8)
where A represents the intercept of the line at K equal to 1 and n represents the slope of
the line [26]. The parameters A and n are material constants. Region III represents high
crack growth rates leading to unstable crack growth and mainly controlled by the fracture
toughness value, K
c
.
Figure 2-35 shows a plot of da/dN versus K for martensitic steels of various
yield and tensile strengths [45]. The tensile strengths for the materials shown in this
figure range from 730 MPa to 1290 MPa. The data fall within a narrow band, with
equations for the upper and lower bands taking the form of Eqn. 5.8 with identical slope
(n = 2.25). The two lines have an average intercept of A = 1.18 x 10
-8
with da/dN in
mm/cycle. As mentioned in the literature review, this plot suggests that the applied stress
intensity factor range is the governing factor controlling crack growth rate behavior rather
than tensile properties (i.e. yield strength, tensile strength and ductility). It should be
noted that the material tested in the current investigation had a tempered martensite
microstructure at all hardness levels and with tensile strengths ranging from 700 MPa to
1525 MPa.
Figure 5-7 shows plots of da/dN versus K for the four hardness levels. The crack
growth rates in Figure 5-7 were calculated using Eqns. 5.3 and 5.5 for crack depth, the a-
N curves presented in Figure 5-5, and the stress intensity factor ranges using Eqn. 5.6.
The stress intensity factor ranges were only calculated for crack depths within the range

159
specified in Eqn. 5.7 (i.e. up to a = 3 mm). It should be noted that the data used in Figure
5-7 are for specimens tested at stress amplitudes similar to the tests at which the crack
length was monitored. Figure 5-7 includes best fit lines to the calculated data points as
well as a prediction line based on the average A and n values from Figure 2-35. The best
fit lines were calculated using crack growth rates above 10
-5
mm/cycle. It can be seen that
the data points calculated are close to the prediction line in all figures.
As previously mentioned, the threshold stress intensity factor range, K
th
, gives
the value at which a crack will not grow. For steels this value has been shown to be much
lower than the fracture toughness value, typically less than 10 m MPa [26]. Figure 5-8
shows crack growth data for all hardness levels superimposed. It can be seen in this
figure that the data deviates from the sloped line in the Paris equation region (Region II)
near crack growth rates below 10
-5
mm/cycle and at a stress intensity factor near
threshold (Region III).
Figure 5-9 shows a schematic plot of Kitagawa-Takahashi diagram giving the
relationship between stress range and crack length. The horizontal line represents the
fatigue limit and the sloped line represents the long crack threshold. The intersection of
the two lines represents the crack length at which short fatigue crack growth ends and
long fatigue crack growth begins and is given by:

2
1
|
|
.
|

\
|
A
A
=
f
th
small
S
K
a
t
(5.9)
where
f
S A is the fatigue limit range. Short or small cracks have an increased crack
growth rate and a lower threshold value than long cracks, as seen in Figure 5-10.

160
Table 5.1 includes calculations of the transition crack depth using Eqn. 5.9. In the
calculation of the transition crack depth, an assumed value of 7 m MPa was used for
th
K A and the fatigue limit range was set equal to twice the fatigue limit from fatigue
tests. It can be see in this table that the calculated transition crack depths are close to
values measured in the metallurgical analysis (0.06 mm to 0.1 mm, see Figure 3-11).
Table 5.1 also includes calculations of the threshold stress intensity factor with Eqn. 5.6
using the fatigue limit from fatigue tests and the crack depth measured from the
metallurgical analysis. The threshold values calculated are around 3 m MPa . This low
value suggests that cracks at the fatigue limit stress level will not grow.

5.2 Fatigue Life Predictions
The following equation was used to estimate the fatigue life of the as-forged
specimens:
n
w
a
n
n
f
f
f A
a a
N
n
n n
)] ( [ ) )( 1 (
2
2 2
2
) 1 (
0
) 1 (
t o A

=

(5.10)
where
f
a is the crack depth at fracture and
0
a is the initial crack depth. Equation 5.10 is
derived by integrating Eqn. 5.8 and assuming a constant geometry factor.
The crack depth at fracture used to estimate fatigue life was calculated using the
following equation:

2
max
)] ( [
1
|
|
.
|

\
|
=
w
a
c
f
f
K
a
o t
(5.11)

161
where
max
o is the maximum bending stress applied to the specimen. In order to solve for
f
a , values of
c
K equal to 65 m MPa for the 45 HRC specimens, 75 m MPa for the 35
HRC specimens, and 80 m MPa for the 25 and 19 HRC specimens were assumed.
These values are similar to the fracture toughness values of other carbon steels with
similar ultimate tensile strength. However, the predicted crack depths became inaccurate
when actual crack depths exceeded the range given by Eqn. 5.7, which occurred at low
stress amplitudes. As a result, fatigue lives were predicted using the actual measured final
crack depths which ranged from 2.5 mm to 5 mm. It should be noted that predictions are
not very sensitive to a
f
(see Figure 5-5), as little life is involved once crack is 2.5 mm
deep.
The initial crack depth was chosen to be 0.06 mm for induction heated forged
specimens and 0.1 mm for gas furnace heated forged specimens. This value came from a
metallurgical analysis performed on the as-forged specimens prior to testing. Figure 3-11
shows an example of surface discontinuities present in the as-forged specimen. The stress
amplitude used to calculate fatigue life was the stress amplitude applied to the fatigue
specimen. A geometry factor of 0.728 for a semi-circular crack, as described in the
previous section, was used for the predictions.
Figures 5-11 and 5-12 show plots of estimated fatigue life versus experimental
fatigue life. These figures have two sets of scatter bands, with the inner bands equal to a
factor of two and the outer bands equal to a factor of three. In Figure 5-11, life for all four
hardness levels was predicted using the average A and n values from Figure 2-35. In
Figure 5-12, life for 45 HRC, 35 HRC, 25 HRC, and 19 HRC specimens was predicted
using the A and n values from the least squares fits from Figure 5-7. Most of the predicted

162
lives in Figures 5-11 and 5-12(a) are within a factor of two or three. The predictions for
the 25 HRC and 19 HRC specimens shown in Figure 5-12(b) are higher than the
predictions for the 45 HRC and 35 HRC specimens. This is due the steeper slope of the
line fits (in Figure 5-7) which results in a lower crack growth rate for a given stress
intensity factor range, therefore longer life. The differences observed between 45 HRC
and 35 HRC specimens compared to 25 HRC and 19 HRC specimens in Figure 5-12 can
be attributed to increased plasticity at the crack tip for the lower hardness material. It can
also be seen in Figure 5-11 that as experimental cycles to failure increases, the estimated
cycles to failure becomes less accurate. This can be explained examining Figure 5-10.
The equation used to predict fatigue life is based on the Paris equation. If the Paris
equation is extrapolated beyond 10
-8
m/cycle (see Figure 5-10), crack growth rates will be
higher and predicted life will, therefore, be shorter.
The effect of the assumed
c
K values used to estimate the final crack depth was
evaluated by predicting life using extreme values of
c
K for 45 HRC specimens. The
results showed small effect on the fatigue life predictions. Figure 5-13 shows a plot of
predicted fatigue life versus experimental fatigue life using
c
K values equal to
30 m MPa , 65 m MPa and 100 m MPa and average A and n values from Figure 2-35.
Using a lower value of
c
K predicts a shorter crack and a shorter life, while larger value
predicts a longer crack and a longer life. However, the difference between the two
predicted lives is small. The reason the final crack depth has little effect on life prediction
is that most of the life is spent when the crack is short.



163
Table 5.1: Summary of threshold stress intensity factor range and transition crack
depth calculations for as-forged specimens forged at different hardness levels.

















Hardness
(HRC)
Heating
Method
Measured
Initial Crack
Depth (mm)
Experimental
Fatigue Limit
(MPa)
a
small
Calculated
Assuming
K
th
(mm)
K
th
Calculated
Assuming a
0
(MPa*m^0.5)
45 Induction 0.06 230 0.07 2.3
45 Gas 0.10 258 0.06 3.3
35 Induction 0.06 303 0.04 3.0
35 Gas 0.10 230 0.07 3.0
25 Induction 0.06 280 0.05 2.8
25 Gas 0.10 225 0.08 2.9
19 Induction 0.06 249 0.06 2.5
19 Gas 0.10 225 0.08 2.9

164
0.1
1
10
0.001 0.01 0.1 1 10
Change in Displacement Amplitude (mm)
M
e
a
s
u
r
e
d

C
r
a
c
k

L
e
n
g
t
h

(
m
m
)























.
45 HRC
35 HRC
25 HRC
19 HRC
R
2
= 0.86
R
2
= 0.97

(a)
0.1
1
10
0.001 0.01 0.1 1 10
Change in Displacement Amplitude (mm)
C
a
l
c
u
l
a
t
e
d

C
r
a
c
k

D
e
p
t
h

(
m
m
)























.
45 HRC
35 HRC
25 HRC
19 HRC
R
2
= 0.86
R
2
= 0.97

(b)

Figure 5-1: Plots of (a) measured crack length, and (b) calculated crack depth versus
change in displacement amplitude for as-forged cantilever bending
specimens at 45 HRC, 35 HRC, 25 HRC and 19 HRC.

165
2
2.5
3
3.5
4
4.5
4 5 6 7 8 9
Measured Crack Length (mm)
M
e
a
s
u
r
e
d

C
r
a
c
k

D
e
p
t
h

(
m
m
)


















.
45 HRC
35 HRC
25 HRC
19 HRC
Least Squares Fit
R
2
= 0.67

Figure 5-2: Measured crack depth versus measured crack length at fracture for 45
HRC, 35 HRC, 25 HRC and 19 HRC specimens having an as-forged
surface from cantilever bending fatigue tests.


(a) (b)
Figure 5-3: Schematic illustrations for as-forged surface specimens plotted in Figure
5.2 of (a) crack depth, a, and crack length, b, and (b) crack shape.

b
a

166







0
.
0
0
.
2
0
.
4
0
.
6
0
.
8
1
.
0
1
.
2
0
.
0
0
.
5
1
.
0
N
o
r
m
a
l
i
z
e
d

C
y
c
l
e
s

N
/
N
f
a ( m m )
0
.
0
0
.
2
0
.
4
0
.
6
0
.
8
1
.
0
1
.
2
0
.
0
0
.
5
1
.
0
N
o
r
m
a
l
i
z
e
d

C
y
c
l
e
s

N
/
N
f
a ( m m )
0
.
0
0
.
2
0
.
4
0
.
6
0
.
8
1
.
0
1
.
2
1
.
4
1
.
6
0
.
0
0
.
5
1
.
0
N
o
r
m
a
l
i
z
e
d

C
y
c
l
e
s

N
/
N
f
a ( m m ) .
0
.
0
0
.
2
0
.
4
0
.
6
0
.
8
1
.
0
1
.
2
1
.
4
1
.
6
1
.
8
2
.
0
0
.
0
0
.
5
1
.
0
N
o
r
m
a
l
i
z
e
d

C
y
c
l
e
s

N
/
N
f

a
( m m )



(
a
)








(
b
)






















(
c
)













(
d
)

F
i
g
u
r
e

5
-
4
:




C
h
a
n
g
e

i
n

d
i
s
p
l
a
c
e
m
e
n
t

a
m
p
l
i
t
u
d
e

v
e
r
s
u
s

n
o
r
m
a
l
i
z
e
d

c
y
c
l
e
s

i
n

c
a
n
t
i
l
e
v
e
r

b
e
n
d
i
n
g

f
a
t
i
g
u
e

t
e
s
t
s

o
f

a
s
-
f
o
r
g
e
d




s
u
r
f
a
c
e

s
p
e
c
i
m
e
n
s

a
t

(
a
)

4
5

H
R
C

h
a
r
d
n
e
s
s

l
e
v
e
l
,

(
b
)

3
5

H
R
C

h
a
r
d
n
e
s
s

l
e
v
e
l
,

(
c
)

2
5

H
R
C

h
a
r
d
n
e
s
s

l
e
v
e
l
,

a
n
d

(
d
)




1
9

H
R
C

h
a
r
d
n
e
s
s

l
e
v
e
l
.



167






0
.
0
0
.
5
1
.
0
1
.
5
2
.
0
2
.
5
3
.
0
3
.
5
0
.
0
0
.
5
1
.
0
N
o
r
m
a
l
i
z
e
d

C
y
c
l
e
s

N
/
N
f
C a l c u l a t e d C r a c k D e p t h ( m m )
0
.
0
0
.
5
1
.
0
1
.
5
2
.
0
2
.
5
3
.
0
3
.
5
4
.
0
0
.
0
0
.
5
1
.
0
N
o
r
m
a
l
i
z
e
d

C
y
c
l
e
s

N
/
N
f
C a l c u l a t e d C r a c k D e p t h ( m m ) .
0
.
0
0
.
5
1
.
0
1
.
5
2
.
0
2
.
5
3
.
0
3
.
5
4
.
0
0
.
0
0
.
5
1
.
0
N
o
r
m
a
l
i
z
e
d

C
y
c
l
e
s

N
/
N
f
C a l c u l a t e d C r a c k D e p t h ( m m ) .
0
.
0
0
.
5
1
.
0
1
.
5
2
.
0
2
.
5
3
.
0
3
.
5
4
.
0
0
.
0
0
.
5
1
.
0
N
o
r
m
a
l
i
z
e
d

C
y
c
l
e
s

N
/
N
f
C a l c u l a t e d C r a c k D e p t h ( m m ) .



(
a
)








(
b
)






















(
c
)













(
d
)

F
i
g
u
r
e

5
-
5
:




C
a
l
c
u
l
a
t
e
d

c
r
a
c
k

d
e
p
t
h

(
u
s
i
n
g

E
q
n
.

5
.
3
)

v
e
r
s
u
s

n
o
r
m
a
l
i
z
e
d

c
y
c
l
e
s

i
n

c
a
n
t
i
l
e
v
e
r

b
e
n
d
i
n
g

f
a
t
i
g
u
e

t
e
s
t
s

o
f

a
s
-
f
o
r
g
e
d




s
u
r
f
a
c
e

s
p
e
c
i
m
e
n
s

a
t

(
a
)

4
5

H
R
C

h
a
r
d
n
e
s
s

l
e
v
e
l
,

(
b
)

3
5

H
R
C

h
a
r
d
n
e
s
s

l
e
v
e
l
,

(
c
)

2
5

H
R
C

h
a
r
d
n
e
s
s

l
e
v
e
l
,

a
n
d

(
d
)




1
9

H
R
C

h
a
r
d
n
e
s
s

l
e
v
e
l
.



168





Figure 5-6: Schematic plot of crack growth rate versus stress intensity factor range
[26].














169

1
.
E
-
0
6
1
.
E
-
0
5
1
.
E
-
0
4
1
.
E
-
0
3
1
.
E
-
0
2
1
1
0
1
0
0

K

(
M
P
a
*
m
^
0
.
5
)
d a / d N ( m m / c y c l e ) .
1
.
E
-
0
6
1
.
E
-
0
5
1
.
E
-
0
4
1
.
E
-
0
3
1
.
E
-
0
2
1
1
0
1
0
0

K

(
M
P
a
*
m
^
0
.
5
)
d a / d N ( m m / c y c l e ) .
1
.
E
-
0
6
1
.
E
-
0
5
1
.
E
-
0
4
1
.
E
-
0
3
1
.
E
-
0
2
1
1
0
1
0
0

K

(
M
P
a
*
m
^
0
.
5
)
d a / d N ( m m / c y c l e ) .
1
.
E
-
0
6
1
.
E
-
0
5
1
.
E
-
0
4
1
.
E
-
0
3
1
.
E
-
0
2
1
1
0
1
0
0

K

(
M
P
a
*
m
^
0
.
5
)
d a / d N ( m m / c y c l e ) .










(
a
)




(
b
)











(
c
)





(
d
)

L
e
g
e
n
d
I
n
d
u
c
t
i
o
n

H
e
a
t
e
d
G
a
s

F
u
r
n
a
c
e

H
e
a
t
e
d
E
s
t
i
m
a
t
i
o
n

M
a
r
t
e
n
s
i
t
i
c
L
e
a
s
t

S
q
u
a
r
e
s

F
i
t
F
i
g
u
r
e

5
-
7
:



C
r
a
c
k

g
r
o
w
t
h

r
a
t
e

(
d
a
/
d
N
)

v
e
r
s
u
s

s
t
r
e
s
s

i
n
t
e
n
s
i
t
y

f
a
c
t
o
r

r
a
n
g
e

(

K
)

f
o
r

f
o
r
g
e
d

s
p
e
c
i
m
e
n
s

s
u
b
j
e
c
t
e
d

t
o

c
a
n
t
i
l
e
v
e
r




b
e
n
d
i
n
g

f
a
t
i
g
u
e

a
t

(
a
)

4
5

H
R
C
,

(
b
)

3
5

H
R
C
,

(
c
)

2
5

H
R
C
,

a
n
d

(
d
)

1
9

H
R
C
.



170
1.E-06
1.E-05
1.E-04
1.E-03
1.E-02
1 10 100
K (MPa*m^0.5)
d
a
/
d
N

(
m
m
/
c
y
c
l
e
)


.
45 HRC
35 HRC
25 HRC
19 HRC
Estimation Martensitic

Figure 5-8: Composite plot of crack growth rate (da/dN) versus stress intensity factor
range (K) for 45 HRC, 35 HRC, 25 HRC, and 19 HRC forged specimens
subjected to cantilever bending fatigue.


171


Figure 5-9: Schematic plot of a Kitagawa-Takahashi diagram giving the relationship
between stress range and crack length [26].




Figure 5-10: Schematic plot of crack growth rate (da/dN) versus stress intensity factor
range (K) showing effect of short cracks and extrapolation of Paris
equation line [26].

172
1.E+03
1.E+04
1.E+05
1.E+06
1.E+07
1.E+03 1.E+04 1.E+05 1.E+06 1.E+07
Experimental Cycles to Failure
E
s
t
i
m
a
t
e
d

C
y
c
l
e
s

t
o

F
a
i
l
u
r
e









.
45 HRC
35 HRC
25 HRC
19 HRC

Figure 5-11: Estimated cycles to failure versus experimental cycles to failure for as-
forged specimens at 45 HRC, 35 HRC, 25 HRC, and 19 HRC using
average A and n values from [45] in Paris equation.


















173
1.E+03
1.E+04
1.E+05
1.E+06
1.E+07
1.E+03 1.E+04 1.E+05 1.E+06 1.E+07
Experimental Cycles to Failure
E
s
t
i
m
a
t
e
d

C
y
c
l
e
s

t
o

F
a
i
l
u
r
e








.
45 HRC
35 HRC

(a)
1.E+03
1.E+04
1.E+05
1.E+06
1.E+07
1.E+03 1.E+04 1.E+05 1.E+06 1.E+07
Experimental Cycles to Failure
E
s
t
i
m
a
t
e
d

C
y
c
l
e
s

t
o

F
a
i
l
u
r
e








.
25 HRC
19 HRC

(b)

Figure 5-12: Estimated cycles to failure versus experimental cycles to failure for as-
forged specimens using experimental A and n values at (a) 45 HRC and 35
HRC, and (b) 25 HRC and 19 HRC.

174
1.E+03
1.E+04
1.E+05
1.E+06
1.E+07
1.E+03 1.E+04 1.E+05 1.E+06 1.E+07
Experimental Cycles to Failure
E
s
t
i
m
a
t
e
d

C
y
c
l
e
s

t
o

F
a
i
l
u
r
e









.
K=30
K=100
K=65

Figure 5-13: Plot of estimated cycles to failure versus experimental cycles to failure
for 45 HRC forged specimens showing effect of assumed K
c
value in life
predictions.


















175






Chapter 6





Prediction of Fatigue Limit and
Mathematical Representation of
Surface Finish Effects


Test data from this study were used for fatigue limit predictions based on models
discussed in the literature review. The results of the fatigue limit predictions are
compared with the experiment results and discussed in this chapter. This chapter also
discusses surface finish factor as functions of hardness and cycles to failure for the
various conditions tested. A simple equation for surface finish factor is then developed,
which can be used for fatigue life analysis and predictions.

6.1 Roughness-Based Prediction Models for Fatigue Limit
6.1.1 Murakami model
Murakami et al [22-24] developed an equation for calculating fatigue limits for
parts containing defects and inclusions. This model uses a fracture mechanics-based
approach for calculating the fatigue limit. In this model, the inclusion or defect is
considered as a crack. This model predicts the stress at which a crack will not propagate.

176
The following equation is used to predict the fatigue limit of parts containing defects and
inclusions for a stress ratio of R = -1:
( )
( )
6
1
120 HV 43 . 1
area
w
+
= o (6.1)
It should be noted that Eqn. 6.1 can be used for steels as well as other materials such as
copper and aluminum. The square root area parameter represents the area of the defect
projected onto the plane perpendicular to the applied stress. This equation was used for a
range of Vickers hardness (HV) between 70 and 720. Murakami then altered Eqn. 6.1 so
that roughness parameters could be used to determine the fatigue limit. This was done by
developing an equivalent square root area parameter that accounts for both depth and
pitch of roughness as defined below:

3 2
) 2 / ( 47 . 9 ) 2 / ( 51 . 3 ) 2 / ( 97 . 2
2
b a b a b a
b
area
R
~ For 195 . 0 2 / < b a (6.2)
38 . 0
2
~
b
area
R
For 195 . 0 2 / 3 > > b a (6.3)
where a is the roughness parameter defined as R
y
which is the maximum depth of
roughness, and 2b represents the pitch. In this study 2b is represented by the roughness
parameter S
m
, which is defined as the mean spacing between peaks at the mean line. It
should be noted that Eqns. 6.1 through 6.3 are discussed in Chapter 2 (see Figure 2-4 and
Figures 2-15 through 2-19). Table 6.1 lists the a/2b values and square root area
parameters calculated using the roughness measurements taken for the specimens with
the as-forged surface condition, as reported in Table 3.2. This table also lists
experimental and predicted fatigue limits as well as the percent difference between the
two values. Figure 6-1 shows a plot of predicted fatigue limit versus experimental fatigue

177
limit along with 10% scatter bands for as-forged specimens at the four hardness levels. It
can be seen from Table 6.1 and Figure 6-1 that about half of the predicted fatigue limits
are within about 10% of the experimental values. The predictions get further away from
experimental values as the hardness increases. However the detrimental effect of surface
roughness typically increases as hardness increases. It should be noted that the model
does not consider the effect of decarburization or surface micro-cracks present in the as-
forged specimens.

6.1.2 Arola-Ramulu model
The Arola-Ramulu model [25] treats surface roughness as a series of notches, as
discussed in section 2.2.5. In their model they developed an equivalent stress
concentration factor which was used to calculate an equivalent fatigue notch factor. The
equivalent stress concentration factor (
t
K ), equivalent fatigue notch factor (
f
K ) and the
equation used to estimate the fatigue life of the as-forged specimens (Eqn. 6.6) are given
as:
|
|
.
|

\
|
|
|
.
|

\
|
+ =
z
y
a
t
R
R
R
n K

1 (6.4)
) 1 ( 1 + =
t f
K q K (6.5)

f
polished f
a
K
S
S
) (
= (6.6)
The stress concentration factor shown in Eqn. 6.4 is calculated using four roughness
parameters, R
a
, R
y
, R
z
and (defined in section 2.2.5). It should be noted that the radius
of the valley, , is rarely quantified as it is the most difficult of the four parameters to

178
measure. The definition of the notch sensitivity, q, can be found in section 2.2.5. Neuber
had previously proposed a similar model for stress concentration factor given by:

z
t
R
n K + =1 (6.7)
It is suggested by Arola and Williams [25] that = 1 for mechanically processed
surfaces. It should be noted that the value of n in Eqns. 6.4 and 6.7 is equal to two for
tension loads. Fatigue notch factors were calculated by substituting both Eqn. 6.4 and 6.7
into Eqn. 6.5. The fatigue limits were then predicted using Eqn. 6.6 for both models.
Table 6.2 lists the parameters used in Eqns. 6.4 through 6.7 with the exception of
the roughness parameters R
a
, R
y
, and R
z
which are listed in Table 3.2. The results shown
in Table 6.2 indicate that the Neuber model predicts a larger fatigue notch factor than the
Arola-Ramulu model, similar to that presented in [25]. The experimental fatigue notch
factor listed in Table 6.2 is defined as the ratio of the fatigue strength of a smooth part to
the fatigue strength of the notched part. Figure 6-2 shows a plot of predicted fatigue limit
versus experimental fatigue limit along with 10% scatter bands using the Arola-Ramulu
and Neuber prediction models for the as-forged specimens at 45 HRC, 35 HRC, 25 HRC,
and 19 HRC. It can be seen in this figure that the predictions are well outside the 10%
scatter bands. If a fatigue notch factor is calculated using data from the current
investigation, the results range from about 25% to 60% greater than factors obtained
using the Neuber and Arola prediction models. The predicted notch factor gets closer to
the experimental notch factor as hardness decreases. In addition, fatigue limits predicted
using the Neuber and Arola-Ramulu models are 25% to 60% greater than experimental
fatigue limits.

179
Although Murakami [23] and Arola and Williams [25] found the equations for
surface roughness to be accurate in their studies, the predictions are less accurate for data
in this study. Both Murakami and Arola-Ramulu prediction models predicted values
closer to experimental data in the current investigation as hardness decreased. Data from
the current investigation shows that the surface hardness becomes closer to the core
hardness as hardness levels decrease. Neither model considers the difference in surface
hardness compared to core hardness resulting from decarburization. The fact that the
prediction models are very accurate for specimens without decarburization and surface
cracks (i.e. just surface roughness) and inaccurate for specimens with as-forged surface
defects, supports the idea that surface roughness is not the only contributing factor to the
fatigue behavior of the steel forgings.

6.2 Mathematical Representation of Experimental Data
6.2.1 Surface finish factor as a function of Brinell hardness
The experimental data has been represented as a factor of two variables, Brinell
hardness and cycles to failure. Although the surface finish factor has been represented
primarily as a function of tensile strength historically, there is a well established
relationship between Brinell hardness and tensile strength for steels, given by:
( ) HB 45 . 3 =
u
S MPa (6.8)
As hardness is one of the simplest material properties to measure, the surface
finish effect is represented as a function of hardness rather than tensile strength for
modeling in this work (see Figure 3-17 for plot of tensile strength versus Brinell
hardness).

180
Table 4.1 summarizes the surface finish factors for the as-forged as well as as-
forged and shot cleaned surface specimens subjected to cantilever and rotating bending
fatigue at the four hardness levels (45 HRC, 35 HRC, 25 HRC, and 19 HRC) at cycles to
failure ranging from 10
4
to 10
6
. Figures 6-3 and 6-4 show plots of the surface finish
factor for both heating methods as a function of Brinell hardness for the as-forged surface
and the shot cleaned surface, respectively. Each figure has four lines representing four
different cycles to failure. The lines represent a linear least squares fit to the data for a
given number of cycles to failure. It should be noted that the surface finish factor was the
dependent variable and hardness was the independent variable. It can be seen that there is
a near linear relationship between surface finish factor and hardness in most cases. The
R
2
values range from 0.99 to 0.72. These figures show that the surface finish factor
increases (i.e. less surface finish effect) with decreasing number of cycles to failure and
as hardness decreases. This is expected due to decreased stress concentration effect at
shorter lives and for more ductile materials (i.e. lower hardness).
Figure 6-5 shows a plot of surface finish factor versus hardness for both induction
heated and gas furnace heated forged specimens with as-forged and as-forged and shot
cleaned surfaces. The lines shown in this figure are fit to data for the four surface
conditions (induction heated as-forged, induction heated as-forged and shot cleaned, gas
furnace heated as-forged, and gas furnace heated as-forged and shot cleaned) for a given
number of cycles to failure. It can be seen in this figure that there are minor differences
between the surface finish factors for the different surface conditions at each hardness
level. The equations for these lines are given by:
082 . 1 ) HB ( 0009 . 0 + =
s
k For 3x10
4
cycles (R
2
= 0.85) (6.9)

181
991 . 0 ) HB ( 0009 . 0 + =
s
k For 10
5
cycles (R
2
= 0.76) (6.10)
023 . 1 ) HB ( 0013 . 0 + =
s
k For 3x10
5
cycles (R
2
= 0.81) (6.11)
921 . 0 ) HB ( 0014 . 0 + =
s
k For 10
6
cycles (R
2
= 0.84) (6.12)

6.2.2 Surface finish factor as a function of cycles to failure
Figures 6-6 and 6-7 show plots of the surface finish factor as a function of cycles
to failure in log scale with both heating methods for the as-forged surface and for the shot
cleaned surface, respectively. There are four lines in each plot, one line for each hardness
level. The lines represent a least squares fit to the data in a semi-log plot with the surface
finish factor as the independent variable. The R
2
values ranged from 0.86 to 0.99. It can
be seen is these figures that the surface finish factor increases for a given hardness as
cycles to failure decreases.
The surface finish factor for a given hardness is calculated by dividing the fatigue
strength of the as-forged surface finish part by the fatigue strength of the smooth and
polished surface part at long life (typically at 10
6
cycles to failure for steels). The surface
finish factors in Table 4.1 for historical data obtained from [2] for life less than 10
6
cycles
were calculated using S-N lines for polished specimens and as-forged specimens that
intersect at 90% of the ultimate tensile strength at 10
3
cycles and extend to a fatigue limit
at 10
6
cycles, as recommended in [13]. The surface finish factors for experimental data
were obtained by dividing the fatigue strength given by the S-N line for the as-forged
surface specimens by the fatigue strength given by S-N line for the machined and
polished surface specimens. The surface finish factors for the current investigation show

182
relationship with cycles to failure similar to historical data, although with a less severe
surface finish effect (see Figures 4-11 through 4-14).
Figure 6-8 shows a plot of surface finish factor versus cycles to failure for both
induction heated and gas furnace heated specimens with as-forged and as-forged and shot
cleaned surfaces. The lines shown in this figure are fit to data for the four surface
conditions (induction heated as-forged, induction heated as-forged and shot cleaned, gas
furnace heated as-forged, and gas furnace heated as-forged and shot cleaned) for a given
hardness level. It can be seen in this figure that there are minor differences between the
surface finish factors for the different surface conditions at four fatigue lives. The
equations for these lines are given by:
83 . 1 ) ( log 205 . 0 + =
f s
N k For 19 HRC (220 HB) (R
2
= 0.86) (6.13)
93 . 1 ) ( log 231 . 0 + =
f s
N k For 25 HRC (253 HB) (R
2
= 0.97) (6.14)
98 . 1 ) ( log 257 . 0 + =
f s
N k For 35 HRC (327 HB) (R
2
= 0.84) (6.15)
04 . 2 ) ( log 289 . 0 + =
f s
N k For 45 HRC (421 HB) (R
2
= 0.92) (6.16)

6.2.3 Surface finish factor as a function of both Brinell hardness and cycles to failure
Figure 6-9(a) shows the surface finish factor as a function of both hardness and
cycles to failure for as-forged specimens with induction heating for hardness ranging
from 220 HB to 421 HB and cycles ranging from 3x10
4
to 10
6
. A surface was fit to all
data points shown in the figure with an R
2
value of 0.97. The equation for the surface has
the form:
( ) ( )
f s
N c b a k log HB = (6.17)

183
The hardness has a linear relationship with the surface finish factor and the cycles to
failure has a logarithmic relationship with the surface finish factor.
Table 6.3 lists the R
2
values using different relationships of hardness to the
surface finish factor (linear, 2
nd
order polynomial, logarithmic, and power functions),
while keeping a logarithmic relationship with cycles to failure. These relations are listed
on the bottom of Table 6.3. It can be seen that there is not much difference in the R
2

values for linear, 2
nd
order polynomial, and logarithmic fits (average R
2
of 0.94 to 0.95).
As a result, the linear relationship was chosen for its simplicity. The values of the
coefficients in Eqn. 6.17 for different surface conditions are given in Table 6.4.
Figure 6-9(b) shows the surface finish factor as a function of both hardness and
cycles to failure for as-forged specimens with gas furnace heating for hardness ranging
from 220 HB to 421 HB and cycles ranging from 3x10
4
to 10
6
. A surface was fit to all
data points shown in the figure with an R
2
value of 0.99. Figure 6-10 shows similar plots
for as-forged and shot cleaned specimens for induction heating (R
2
= 0.93) and gas
furnace heating (R
2
= 0.95).
It can be seen that there is not much difference in the equation constants a, b, and
c for the surfaces fit to data from each heating method. As a result, a surface was fit to as-
forged surface data from both heating methods for hardness ranging from 220 HB to 421
HB and cycles ranging from 3x10
4
to 10
6
, shown in Figure 6-11(a), with an R
2
value of
0.95. It should be noted that the surface finish factors shown in this figure for the
induction heated forged specimens at 45 HRC are slightly lower than for the gas furnace
heated forged specimens at all cycles to failure. The equation for the surface shown in
Figure 6-11(a) is given by:

184
( ) ( )
f s
N k log 232 . 0 HB 00135 . 0 26 . 2 = (6.18)
Substituting Eqn. 6.8 into Eqn. 6.18 gives a relationship between surface finish factor,
ultimate tensile strength (in MPa), and cycles to failure for the as-forged surface
condition as follows:
( ) ( )
f u s
N S k log 232 . 0 000391 . 0 26 . 2 = (6.19)
In Figure 6-11(b), similar to Figure 6-11(a), the surface finish factors for the induction
heated forged and shot cleaned specimens at 45 HRC are slightly lower than for the gas
furnace heated forged and shot cleaned specimens at all cycles to failure. The equation
for the surface (R
2
= 0.91) shown in Figure 6-11(b) is given by:
( ) ( )
f s
N k log 212 . 0 HB 00134 . 0 21 . 2 = (6.20)
Substituting Eqn. 6.8 into Eqn. 6.20 gives the relationship in terms of ultimate tensile
strength as:
( ) ( )
f u s
N S k log 212 . 0 000388 . 0 21 . 2 = (6.21)
Figure 6-12 shows a plot of surface finish factor versus Brinell hardness and
cycles to failure with a surface fit to all four surface conditions (induction heated as-
forged, induction heated as-forged and shot cleaned, gas furnace heated as-forged, and
gas furnace heated as-forged and shot cleaned). The equation for the surface (R
2
= 0.88)
shown in Figure 6-12 is given by:
( ) ( )
f s
N k log 222 . 0 HB 00134 . 0 24 . 2 = (6.22)
In terms of ultimate tensile strength, the relation is:
( ) ( )
f u s
N S k log 222 . 0 000388 . 0 24 . 2 = (6.23)

185
It can be seen in Figure 6-12 that the surface finish factors for the two heating methods
are somewhat closer to each other than the surface finish factors for the shot cleaned and
as-forged (not shot cleaned) conditions of each heating method.
Although the surface finish factor was only fit to a range of hardness between 220
HB and 421 HB and a range of cycles between 3x10
4
and 10
6
,

it is reasonable to use Eqn.
6.17 for cycles to failure ranging from 10
4
to 10
6
cycles and for hardness values ranging
from 150 HB to 500 HB. Figure 6-13(a) shows a plot of the experimental surface finish
factor for the as-forged surface finish specimens versus the surface finish factor
calculated using Eqn. 6.18. The factors based on the historical data from [2] are also
included, see section 6.2.2 for a description of how surface finish factor (k
s
) values were
calculated. Data points plotted in this figure were calculated for hardness ranging from
220 HB to 421 HB and cycles ranging from 10
4
to 10
6
. This figure includes a 45 line as
well as 5% and 10% scatter bands. It can be seen that most of the experimental surface
finish factors fall within the 5% bands and the rest falling in the 10% bands. However the
old surface finish factors are significantly outside the 10% bands (see Figure 6-13(a)). It
should be noted that there is little to no surface finish effect on fatigue behavior for cycles
to failure less than 10
4
, as seen in Figure 6-13(a). Figure 6-13(b) shows a similar plot for
as-forged and shot cleaned surface specimens using Eqn. 6.20. Most of the experimental
surface finish factors fall within the 5% and 10% bands in this figure as well.
Figure 6-14(a) shows a plot of calculated surface finish factor versus Brinell
hardness for cycles to failure ranging from 10
4
to 10
6
cycles. The surface finish factor in
Figure 6-14(a) was calculated using Eqn. 6.18. Figure 6-14(b) shows a similar plot,

186
except using Eqn. 6.20 to calculate the surface finish factor. Such plots can serve as
surface finish factor charts for design applications.










































187
Table 6.1: Summary of calculations of square root area parameters and fatigue limit
predictions using Murakami model for as-forged specimens.

















(HRC) (HV)
Induction 0.048 108 371 231 1.61
Induction-shot
cleaned
0.155 104 373 269 1.39
Gas furnace 0.050 88 384 258 1.49
Gas furnace-
shot cleaned
0.121 70 399 310 1.28
Induction 0.165 95 311 303 1.03
Induction-shot
cleaned
0.126 115 301 361 0.84
Gas furnace 0.174 84 318 230 1.38
Gas furnace-
shot cleaned
0.234 75 324 287 1.13
Induction 0.149 116 250 280 0.89
Induction-shot
cleaned
0.099 127 246 305 0.81
Gas furnace 0.161 110 252 225 1.12
Gas furnace-
shot cleaned
0.127 82 265 289 0.92
Induction 0.147 111 232 249 0.93
Induction-shot
cleaned
0.169 103 234 312 0.75
Gas furnace 0.119 84 243 225 1.08
Gas furnace-
shot cleaned
0.182 101 235 283 0.83
19 235
35 345
25 266
Fatigue
Limit Exp.
(MPa)

pred.
/
exp.
45 446
Surface
Condition
a/2b
(R
y
/S
m
)
Square
root area
parameter
(m)
Fatigue
Limit
Murakami
(MPa)
Hardness

188


T
a
b
l
e

6
.
2
:




S
u
m
m
a
r
y

o
f

c
a
l
c
u
l
a
t
i
o
n
s

o
f

s
t
r
e
s
s

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

f
a
c
t
o
r
s
,

f
a
t
i
g
u
e

n
o
t
c
h

f
a
c
t
o
r
s
,

a
n
d

f
a
t
i
g
u
e

l
i
m
i
t

p
r
e
d
i
c
t
i
o
n
s

u
s
i
n
g




A
r
o
l
a
-
R
a
m
u
l
u

a
n
d

N
e
u
b
e
r

m
o
d
e
l
s

f
o
r

a
s
-
f
o
r
g
e
d

s
p
e
c
i
m
e
n
s
.



189
Table 6.3: Summary of R
2
values used in three dimensional fits to the surface finish
factor data as a function of hardness and cycles to failure for the different
surface conditions tested.



Table 6.4: Summary of constants for use in Eqn. 6.17 for the three dimensional
surface finish data fits shown in Figures 6-9 through 6-12.


a b c
Induction As-forged 2.35 -0.00155 -0.234
Induction Shot
Cleaned
2.27 -0.00155 -0.208
Gas Furnace As-
forged
2.18 -0.00114 -0.230
Gas Furnace Shot
Cleaned
2.15 -0.00113 -0.216
Induction and Gas
Furnace As-forged
2.26 -0.00135 -0.232
Induction and Gas
Furnace Shot
Cleaned
2.21 -0.00134 -0.212
Induction and Gas
Furnace As-Forged
and Shot Cleaned
2.24 -0.00134 -0.222
Surface Condition
Constants
Eqn. 1 Eqn. 2 Eqn. 3 Eqn. 4
Induction As-
forged
0.965 0.967 0.940 -
Induction
Shot Cleaned
0.932 0.935 0.901 -
Gas Furnace
As-forged
0.988 0.982 0.984 0.984
Gas Furnace
Shot Cleaned
0.951 0.961 0.960 0.918
Induction and
Gas Furnace
As-forged
0.954 0.953 0.954 0.596
Induction and
Gas Furnace
Shot Cleaned
0.910 0.909 0.908 0.566
Eqn. 1: k
s
= a + b (HB) + c log(N
f
)
Eqn. 2: k
s
= a + b (HB)
2
+ c (HB) + d log(N
f
)
Eqn. 3: k
s
= a + b log(HB) + c log(N
f
)
Eqn. 4: k
s
= a + b (HB)
c
+ d log(N
f
)
Surface
Condition
R
2

190
0
50
100
150
200
250
300
350
400
450
500
550
600
650
0 50 100 150 200 250 300 350 400 450 500 550 600 650
Experimental Fatigue Limit (MPa)
P
r
e
d
i
c
t
e
d

F
a
t
i
g
u
e

L
i
m
i
t

(
M
P
a
)


.
As- Forged 45 HRC
As-Forged 35 HRC
As-Forged 25 HRC
As-Forged 19 HRC
Shot Cleaned 45 HRC
Shot Cleaned 35 HRC
Shot Cleaned 25 HRC
Shot Cleaned 19 HRC


Figure 6-1: Predicted fatigue limit versus experimental fatigue limit using Murakami
model for as-forged specimens at 45 HRC, 35 HRC, 25 HRC, and 19
HRC.









191
0
100
200
300
400
500
600
700
0 100 200 300 400 500 600 700
Experimental Fatigue Limit (MPa)
P
r
e
d
i
c
t
e
d

F
a
t
i
g
u
e

L
i
m
i
t

(
M
P
a
)


.
As- Forged 45 HRC
As-Forged 35 HRC
As-Forged 25 HRC
As-Forged 19 HRC
Shot Cleaned 45 HRC
Shot Cleaned 35 HRC
Shot Cleaned 25 HRC
Shot Cleaned 19 HRC

(a)
0
50
100
150
200
250
300
350
400
450
500
550
600
650
0 50 100 150 200 250 300 350 400 450 500 550 600 650
Experimental Fatigue Limit (MPa)
P
r
e
d
i
c
t
e
d

F
a
t
i
g
u
e

L
i
m
i
t

(
M
P
a
)


.
As- Forged 45 HRC
As-Forged 35 HRC
As-Forged 25 HRC
As-Forged 19 HRC
Shot Cleaned 45 HRC
Shot Cleaned 35 HRC
Shot Cleaned 25 HRC
Shot Cleaned 19 HRC

(b)

Figure 6-2: Predicted fatigue limit versus experimental fatigue limit for as-forged
specimens at 45 HRC, 35 HRC, 25 HRC, and 19 HRC using (a) Arola-
Ramulu model, and (b) Neuber model.

192

(a)

(b)

Figure 6-3: Surface finish factor (k
s
) versus hardness at different fatigue lives for as-
forged specimens with (a) induction heating, and (b) gas furnace heating.
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
200 250 300 350 400 450
Hardness, (HB)
10^6
3x10^5
10^5
3x10^4
S
u
r
f
a
c
e

F
i
n
i
s
h

F
a
c
t
o
r
,

k
s
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
200 250 300 350 400 450
Hardness, (HB)
10^6
3x10^5
10^5
3x10^4
S
u
r
f
a
c
e

F
i
n
i
s
h

F
a
c
t
o
r
,

k
s

193

(a)

(b)

Figure 6-4: Surface finish factor (k
s
) versus hardness at different fatigue lives for shot
cleaned specimens with (a) induction heating, and (b) gas furnace heating.



0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
200 250 300 350 400 450
Hardness, (HB)
10^6
3x10^5
10^5
3x10^4
S
u
r
f
a
c
e

F
i
n
i
s
h

F
a
c
t
o
r
,

k
s
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
200 250 300 350 400 450
Hardness, (HB)
10^6
3x10^5
10^5
3x10^4
S
u
r
f
a
c
e

F
i
n
i
s
h

F
a
c
t
o
r
,

k
s

194



0
.
0
0
.
1
0
.
2
0
.
3
0
.
4
0
.
5
0
.
6
0
.
7
0
.
8
0
.
9
1
.
0
2
0
0
2
5
0
3
0
0
3
5
0
4
0
0
4
5
0
H
a
r
d
n
e
s
s
,

(
H
B
)
G
a
s

F
u
r
n
a
c
e

H
e
a
t
i
n
g
I
n
d
u
c
t
i
o
n

H
e
a
t
i
n
g
G
a
s

F
u
r
n
a
c
e

H
e
a
t
i
n
g

S
h
o
t

C
l
e
a
n
e
d
I
n
d
u
c
t
i
o
n

H
e
a
t
i
n
g

S
h
o
t

C
l
e
a
n
e
d
G
a
s

F
u
r
n
a
c
e

H
e
a
t
i
n
g

I
n
d
u
c
t
i
o
n

H
e
a
t
i
n
g
G
a
s

F
u
r
n
a
c
e

H
e
a
t
i
n
g

S
h
o
t

C
l
e
a
n
e
d
I
n
d
u
c
t
i
o
n

H
e
a
t
i
n
g

S
h
o
t

C
l
e
a
n
e
d
G
a
s

F
u
r
n
a
c
e

H
e
a
t
i
n
g
I
n
d
u
c
t
i
o
n

H
e
a
t
i
n
g
G
a
s

F
u
r
n
a
c
e

H
e
a
t
i
n
g

S
h
o
t

C
l
e
a
n
e
d
I
n
d
u
c
t
i
o
n

H
e
a
t
i
n
g

S
h
o
t

C
l
e
a
n
e
d
G
a
s

F
u
r
n
a
c
e

H
e
a
t
i
n
g
I
n
d
u
c
t
i
o
n

H
e
a
t
i
n
g
G
a
s

F
u
r
n
a
c
e

H
e
a
t
i
n
g

S
h
o
t

C
l
e
a
n
e
d
I
n
d
u
c
t
i
o
n

H
e
a
t
i
n
g

S
h
o
t

C
l
e
a
n
e
d
1
0
^
6
3
x
1
0
^
5
1
0
^
5
3
x
1
0
^
4
S u r f a c e F i n i s h F a c t o r , k
s
F
i
g
u
r
e

6
-
5
:



S
u
r
f
a
c
e

f
i
n
i
s
h

f
a
c
t
o
r

v
e
r
s
u
s

B
r
i
n
e
l
l

h
a
r
d
n
e
s
s

f
o
r

a
s
-
f
o
r
g
e
d

a
n
d

a
s
-
f
o
r
g
e
d

a
n
d

s
h
o
t

c
l
e
a
n
e
d

s
p
e
c
i
m
e
n
s

w
i
t
h

g
a
s




f
u
r
n
a
c
e

a
n
d

i
n
d
u
c
t
i
o
n

h
e
a
t
i
n
g

a
t

1
0
6
,

3
x
1
0
5
,

1
0
5
,

a
n
d

3
x
1
0
4

c
y
c
l
e
s

t
o

f
a
i
l
u
r
e
.



195

(a)

(b)

Figure 6-6: Surface finish factor (k
s
) versus cycles to failure at different hardness
levels for as-forged specimens with (a) induction heating, and (b) gas
furnace heating.
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
1.E+04 1.E+05 1.E+06 1.E+07
Cycles to Failure, N
f
19 HRC
25 HRC
35 HRC
45 HRC
S
u
r
f
a
c
e

F
i
n
i
s
h

F
a
c
t
o
r
,

k
s
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
1.E+04 1.E+05 1.E+06 1.E+07
Cycles to Failure, N
f
19 HRC
25 HRC
35 HRC
45 HRC
S
u
r
f
a
c
e

F
i
n
i
s
h

F
a
c
t
o
r
,

k
s

196



(a)

(b)

Figure 6-7: Surface finish factor (k
s
) versus cycles to failure at different hardness
levels for shot cleaned specimens with (a) induction heating, and (b) gas
furnace heating.
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
1.E+04 1.E+05 1.E+06 1.E+07
Cycles to Failure, N
f
19 HRC
25 HRC
35 HRC
45 HRC
S
u
r
f
a
c
e

F
i
n
i
s
h

F
a
c
t
o
r
,

k
s
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
1.E+04 1.E+05 1.E+06 1.E+07
Cycles to Failure, N
f
19 HRC
25 HRC
35 HRC
45 HRC
S
u
r
f
a
c
e

F
i
n
i
s
h

F
a
c
t
o
r
,

k
s

197

0
.
0
0
.
1
0
.
2
0
.
3
0
.
4
0
.
5
0
.
6
0
.
7
0
.
8
0
.
9
1
.
0
1
.
E
+
0
4
1
.
E
+
0
5
1
.
E
+
0
6
1
.
E
+
0
7
C
y
c
l
e
s

t
o

F
a
i
l
u
r
e
,

N
f
G
a
s

F
u
r
n
a
c
e

H
e
a
t
i
n
g
I
n
d
u
c
t
i
o
n

H
e
a
t
i
n
g
G
a
s

F
u
r
n
a
c
e

H
e
a
t
i
n
g

S
h
o
t

C
l
e
a
n
e
d
I
n
d
u
c
t
i
o
n

H
e
a
t
i
n
g

S
h
o
t

C
l
e
a
n
e
d
G
a
s

F
u
r
n
a
c
e

H
e
a
t
i
n
g
I
n
d
u
c
t
i
o
n

H
e
a
t
i
n
g
G
a
s

F
u
r
n
a
c
e

H
e
a
t
i
n
g

S
h
o
t

C
l
e
a
n
e
d
I
n
d
u
c
t
i
o
n

H
e
a
t
i
n
g

S
h
o
t

C
l
e
a
n
e
d
G
a
s

F
u
r
n
a
c
e

H
e
a
t
i
n
g
I
n
d
u
c
t
i
o
n

H
e
a
t
i
n
g
G
a
s

F
u
r
n
a
c
e

H
e
a
t
i
n
g

S
h
o
t

C
l
e
a
n
e
d
I
n
d
u
c
t
i
o
n

H
e
a
t
i
n
g

S
h
o
t

C
l
e
a
n
e
d
G
a
s

F
u
r
n
a
c
e

H
e
a
t
i
n
g
I
n
d
u
c
t
i
o
n

H
e
a
t
i
n
g
G
a
s

F
u
r
n
a
c
e

H
e
a
t
i
n
g

S
h
o
t

C
l
e
a
n
e
d
I
n
d
u
c
t
i
o
n

H
e
a
t
i
n
g

S
h
o
t

C
l
e
a
n
e
d
4
5

H
R
C
3
5

H
R
C
2
5

H
R
C
1
9

H
R
C
S u r f a c e F i n i s h F a c t o r , k
s
F
i
g
u
r
e

6
-
8
:



S
u
r
f
a
c
e

f
i
n
i
s
h

f
a
c
t
o
r

v
e
r
s
u
s

c
y
c
l
e
s

t
o

f
a
i
l
u
r
e

f
o
r

a
s
-
f
o
r
g
e
d

a
n
d

a
s
-
f
o
r
g
e
d

a
n
d

s
h
o
t

c
l
e
a
n
e
d

s
p
e
c
i
m
e
n
s

w
i
t
h

g
a
s




f
u
r
n
a
c
e

a
n
d

i
n
d
u
c
t
i
o
n

h
e
a
t
i
n
g

a
t

4
5

H
R
C

(
4
2
1

H
B
)
,

3
5

H
R
C

(
3
2
7

H
B
)
,

2
5

H
R
C

(
2
5
3

H
B
)
,

a
n
d

1
9

H
R
C




(
2
2
0

H
B
)
.



198
10
5
10
6
250
300
350
400
0
0.2
0.4
0.6
0.8
1
Cycles to Failure
Brinell Hardness
S
u
r
f
a
c
e

F
i
n
i
s
h

F
a
c
t
o
r

(a)

10
5
10
6
250
300
350
400
0
0.2
0.4
0.6
0.8
1
Cycles to Failure
Brinell Hardness
S
u
r
f
a
c
e

F
i
n
i
s
h

F
a
c
t
o
r

(b)

Figure 6-9: 3D plot showing surface finish factor as a function of hardness and cycles
to failure for (a) induction heated forged, and (b) gas furnace heated
forged specimens.


199
10
5
10
6
250
300
350
400
0
0.2
0.4
0.6
0.8
1
Cycles to Failure
Brinell Hardness
S
u
r
f
a
c
e

F
i
n
i
s
h

F
a
c
t
o
r

(a)
10
5
10
6
250
300
350
400
0
0.2
0.4
0.6
0.8
1
Cycles to Failure
Brinell Hardness
S
u
r
f
a
c
e

F
i
n
i
s
h

F
a
c
t
o
r

(b)

Figure 6-10: 3D plot showing surface finish factor as a function of hardness and cycles
to failure for (a) induction heated forged and shot cleaned, and (b) gas
furnace heated forged and shot cleaned specimens.

200

10
5
10
6
250
300
350
400
0
0.2
0.4
0.6
0.8
1

Cycles to Failure
Brinell Hardness
S
u
r
f
a
c
e

F
i
n
i
s
h

F
a
c
t
o
r
Surface
As-Forged Gas
As-Forged Induction

(a)
10
5
10
6
250
300
350
400
0
0.2
0.4
0.6
0.8
1

Cycles to Failure
Brinell Hardness
S
u
r
f
a
c
e

F
i
n
i
s
h

F
a
c
t
o
r
Surface
Shot Cleaned Gas
Shot Cleaned Induction

(b)

Figure 6-11: 3D plot showing surface finish factor as a function of hardness and cycles
to failure for induction heated and gas furnace heated forged specimens
with (a) as-forged surface, and (b) as-forged and shot cleaned surface.

201











1
0
5
1
0
6
2
5
0
3
0
0
3
5
0
4
0
0
0
0
.
1
0
.
2
0
.
3
0
.
4
0
.
5
0
.
6
0
.
7
0
.
8
0
.
9 1

C
y
c
l
e
s

t
o

F
a
i
l
l
u
r
e
B
r
i
n
e
l
l

H
a
r
d
n
e
s
s

S u r f a e F i n i s h F a c t o r
S
u
r
f
a
c
e
A
s
-
F
o
r
g
e
d

G
a
s
A
s
-
F
o
r
g
e
d

I
n
d
u
c
t
i
o
n
S
h
o
t

C
l
e
a
n
e
d

G
a
s
S
h
o
t

C
l
e
a
n
e
d

I
n
d
u
c
t
i
o
n
F
i
g
u
r
e

6
-
1
2
:




3
D

p
l
o
t

s
h
o
w
i
n
g

s
u
r
f
a
c
e

f
i
n
i
s
h

f
a
c
t
o
r

a
s

a

f
u
n
c
t
i
o
n

o
f

h
a
r
d
n
e
s
s

a
n
d

c
y
c
l
e
s

t
o

f
a
i
l
u
r
e

f
o
r

i
n
d
u
c
t
i
o
n

h
e
a
t
e
d

a
n
d

g
a
s




f
u
r
n
a
c
e

h
e
a
t
e
d

f
o
r
g
e
d

s
p
e
c
i
m
e
n
s

w
i
t
h

a
s
-
f
o
r
g
e
d

s
u
r
f
a
c
e

a
n
d

a
s
-
f
o
r
g
e
d

a
n
d

s
h
o
t

c
l
e
a
n
e
d

s
u
r
f
a
c
e
.



202
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
1.1
1.2
0.2 0.4 0.6 0.8 1.0 1.2
Calculated Surface Finish Factor From Fit
E
x
p
e
r
i
m
e
n
t
a
l

S
u
r
f
a
c
e

F
i
n
i
s
h

F
a
c
t
o
r



.
45 HRC (Experimental)
35 HRC (Experimental)
25 HRC (Experimental)
19 HRC (Experimental)
45 HRC (Old Data)
35 HRC (Old Data)
25 HRC (Old Data)
19 HRC (Old Data)

(a)
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
1.1
1.2
0.2 0.4 0.6 0.8 1.0 1.2
Calculated Surface Finish Factor From Fit
E
x
p
e
r
i
m
e
n
t
a
l

S
u
r
f
a
c
e

F
i
n
i
s
h

F
a
c
t
o
r



.
45 HRC (Experimental)
35 HRC (Experimental)
25 HRC (Experimental)
19 HRC (Experimental)
45 HRC (Old Data)
35 HRC (Old Data)
25 HRC (Old Data)
19 HRC (Old Data)

(b)

Figure 6-13: Surface finish factor from experimental data and old data versus surface
finish factor calculated from fit to experimental data (both induction and
gas furnace heated specimens) for failure ranging from 10
4
to 10
6
cycles
for (a) as-forged surface, and (b) as-forged and shot cleaned surface.
Scatter bands of 5% and 10% are shown.

203


(a)

(b)

Figure 6-14: Surface finish factor versus hardness for 10
4
, 10
5
, and 10
6
cycles to failure
for (a) as-forged surface using Eqn. 6.18, and (b) as-forged and shot
cleaned surface using Eqn. 6.20.

0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
100 200 300 400 500 600
Brinell Hardness, (HB)
10^6 Cycles
10^5 Cycles
10^4 Cycles
S
u
r
f
a
c
e

F
i
n
i
s
h

F
a
c
t
o
r
,

k
s
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
100 200 300 400 500 600
Brinell Hardness, (HB)
10^6 Cycles
10^5 Cycles
10^4 Cycles
S
u
r
f
a
c
e

F
i
n
i
s
h

F
a
c
t
o
r
,

k
s

204







Chapter 7





Summary and Conclusions







This study investigated the surface finish effects of as-forged surface on the
fatigue behavior of commonly used forging steel. The purpose of this study was to
develop surface finish factors for a range of hardness levels and cycles to failure. The
surface finish factors obtained from fatigue testing were then compared to surface finish
factors currently available based on historical data. Two surface conditions were
evaluated, a smooth-polished surface finish to be used as the reference surface, and a hot-
forged surface finish. The heating methods used for forging were gas furnace heating as
well as induction heating, to allow comparison of the two heating methods, as surface
roughness and decarburization depths differ between the two methods. Since residual
stresses from shot cleaning play a critical synergistic role with surface finish, the hot-
forged surface finish was evaluated with and without residual compressive stresses. Some
testing was also conducted to investigate the effect of the flash left by the forging
process. In addition, the effect of grain flow resulting from the forging process was
evaluated by testing smooth specimens machined from the same rolled bar used for
forging. Based on several hundred fatigue tests performed, as well as the analysis
conducted the following conclusions can be made.

205
1. There was little to no difference in fatigue behaviors for forged specimens under
rotating bending and cantilever bending loads. In theory, smooth specimens under
rotating bending fatigue should experience shorter life compared to reverse
cantilever bending fatigue due to the high volume subjected to highest stress in
rotating bending. However, due to the presence of the as-forged surface
discontinuities, there was less of an effect of rotating bending versus revered
cantilever bending on the fatigue strength at 10
6
cycles for the as-forged
condition, compared to the polished surface condition.
2. The effect of forging flash was found to be insignificant under both cantilever
bending and rotating bending fatigue. This can be attributed to the additional
material from the flash reducing the stress, as well as due to the surface roughness
and discontinuities which may outweigh any stress concentration effects of the
flash.
3. As hardness decreases, the fatigue strength also decreases for machined and
polished specimens, as expected. However, there was little to no increase in
fatigue strength at 10
6
cycles as the hardness or tensile strength increased for
specimens having as-forged surface condition. This is due to surface
decarburization and surface roughness, as harder materials are more sensitive to
stress concentrations.
4. Forging using induction heating resulted in less decarburization than gas furnace
heating. The gas furnace heated forged specimens also had surface discontinuities
almost twice the depth of the induction heated forged specimens. Although the
induction heated specimens had higher surface roughness they still exhibited

206
somewhat better fatigue behavior at all hardness levels, except at 45 HRC,
compared to the gas furnace heated specimens. As a result, the main cause of
difference in fatigue behavior between the two heating methods is believed to be
due to decarburization and surface discontinuities.
5. Shot cleaning improved fatigue behavior in the HCF region due to the surface
compressive residual stress it imparts on the part. Shot cleaning resulted in 8% to
17% increase in fatigue strength at 10
6
cycles, depending on hardness level. The
increase in fatigue strength was similar between gas furnace heated and induction
heated specimens. The beneficial effect was greater at higher hardness levels and
at longer life.
6. There was little or no effect of grain flow observed on fatigue life in this study.
The specimen geometry used with smooth gage section geometry is not suitable to
demonstrate possible beneficial effects of grain flow of a forged part, compared to
a machined part where the machining process can cut across the grains (for
example, a forged stepped shaft versus machined stepped shaft).
7. Forged surface correction factors based on historical data are significantly lower
than the factors obtained in this investigation, therefore very conservative. Use of
the old correction factors can result in predicted fatigue lives in high cycle fatigue
which are significantly shorter than fatigue lives observed from the current
investigation. Both the current as well as historical data suggest that higher
hardness materials exhibit more of a difference in high cycle fatigue strength
between the as-forged and machined and polished surface conditions.

207
8. Fracture mechanics analysis provided fairly accurate representation of the fatigue
data with fatigue life predictions mostly within a factor of two or three for all four
hardness levels tested. This is because there is little difference in crack growth
behavior for steels of a wide range of tensile strengths.
9. An equation was developed to represent the surface finish factor as a function of
both Brinell hardness (or ultimate tensile strength) and cycles to failure. This
equation was shown to result in a surface finish factor within 10% of the
experimental values for the four hardness levels tested and life cycles ranging
from 10
4
to 10
6
.






























208





References






[1] G. H. Farrahi, D. J. Smith, W. X. Zhu, and C. A. McMahon, 2002, Influence of
residual stress on fatigue life of hot forged and shot blasted steel components,
International Journal of Engineering Transactions B: Applications, Vol. 15, pp.
79-86.

[2] G. Noll and C. Lipson, 1946, Allowable working stresses, Proc. Society for
Experimental Analysis, Vol. 3, No. 2, pp. 89-109.

[3] G. Hankins and M. Becker, 1932, The fatigue resistance of unmachined forged
steels, Journal of Iron and Steel Institute, Vol. 126, pp. 205-236.

[4] G. Hankins, M. Becker, and H. Mills, 1936, Further experiments on the effect of
surface finish conditions on fatigue resistance of steels, Journal of the Iron and
Steel Institute, Vol. 133, pp. 309-425.

[5] I. Shareef and M. D. Hasselbusch, 1996, Endurance limit modifying factors for
hardened machined surfaces, SAE Transactions, Vol. 105, No. 5, pp. 889-899.

[6] J. Davies and P. Simpson, 1979, Induction Heating Handbook, McGraw-Hill,
London, NY.

[7] V. Rudnev, D. Loveless, R. Cook, and M. Black, 2002, Handbook of Induction
Heating, Marcel Dekker, New York, NY.

[8] T. Altan, G. Ngaile, and G. Shen, 2005, Cold and Hot Forging Fundamentals and
Applications, ASM International, Materials Park, OH.

[9] T. Altan, S. Oh, and H. Gegel, 1983, Metal Forming: Fundamentals and
Applications, ASM International, Materials Park, OH.

[10] W. Naujoks, 1953, Forging, ASME Handbook: Metals Engineering Design,
Vol. 4, pp. 66-74, McGraw-Hill, New York, NY.

[11] S. Collins and G. Michal, 1995, Forging effects on fatigue properties of AISI
4140 steel, 36
th
MWSP Conf. Proc., Vol. 32, pp. 257-269, ISS-AIME, Baltimore,
MD.


209
[12] Y. Chastel, N. Caillet, and P. Bouchard, 2006, Qualitative analysis of the impact
of forging operations on fatigue properties of steel components, Journal of
Materials Processing Technology, Vol. 177, pp. 202-205.

[13] C. Lipson and G. Noll, 1953, Design practice, ASME Handbook: Metals
Engineering Design, Vol. 3, pp. 297-305, McGraw-Hill, New York, NY.

[14] M. Gildersleeve, 1991, Relationship between decarburization and fatigue
strength of through hardened and carburizing steels, Material Science and
Technology, Vol. 7, pp. 307-310.

[15] K. Adamaszek and P. Broz, 2001, Decarburization and hardness changes in
carbon steels caused by high temperature surface oxidation in ambient air,
Diffusion and Defect Data: Defect and Diffusion Forum, Vol. 194, pp. 1701-
1706.

[16] L. Leonard and V. Toal, 1998, Roughness measurements of metallic surfaces
based on the laser speckle contrast method, Optics and Lasers Engineering, Vol.
30, pp. 433-440.

[17] E. Abbott, 1940, The tracer method of measuring surface irregularities, Surface
Treatments, Vol. 28, pp. 392-427.

[18] E. S. Gadelmawla, M. M. Koura, T. M. A. Maksoud, I. M. Elewa, and H. H.
Soliman, 2002, Roughness parameters, Journal of Materials Processing
Technology, Vol. 123, pp. 133-145.

[19] D. Novovic, R. C. Dewes, D. K. Aspinwall, W. Voice, and P. Bowen, 2004,
Effect of machined topography and integrity on fatigue life, International
Journal of Machine Tools and Manufacture, Vol. 44, pp. 125-134.

[20] P. Fluck, 1951, Influence of surface roughness on the fatigue life and scatter of
test results of two steels, Proceedings of American Society for Testing and
Materials, Vol. 51, pp. 584-592, ASTM, Philadelphia, PA.

[21] E. Siebel and M. Gaier, 1957, The influence of surface roughness on the fatigue
strength of steels and non-ferrous alloys, The Engineers Digest, Vol. 18, No. 3,
pp. 109-112.

[22] Y. Murakami and M. Endo, 1994, Effect of defects, inclusions and
inhomogenities on fatigue strength, International Journal of Fatigue, Vol.16,
No. 3, pp. 163-182.

[23] Y. Murakami, 2002, Effect of surface roughness on fatigue strength, Metal
Fatigue: Effect of Small Defects and Non Metallic Inclusions, pp. 28-40, Elsevier,
Kidlington, Oxford, UK.

210

[24] K. Takahashi and Y. Murakami, 1999, Quantitative evaluation of effect of
surface roughness on fatigue strength, Engineering Against Fatigue, pp. 693-
703, A.A. Balkema, Ed., Sheffield, UK.

[25] D. Arola and C. L. Williams, 2002, Estimating the fatigue stress concentration
factor of machined surfaces, International Journal of Fatigue, Vol. 24, pp. 923-
930.

[26] R. I. Stephens, A. Fatemi, R. R. Stephens, and H. O. Fuchs, 2000, Metal Fatigue
in Engineering, 2
nd
Ed., John Wiley and Sons, New York, NY.

[27] R. Budynas and J. Nisbett, 2008, Shigleys Mechanical Engineering Design, 8
th

Ed., McGraw Hill, New York, NY.

[28] C. Lipson and R. Juvinall, 1963, Handbook of Stress and Strength, pp. 99-113,
Macmillan, New York, NY.

[29] K. Edwards and R. McKee, 1991, Fundamentals of Mechanical Component
Design, McGraw-Hill, New York, NY.

[30] R. Norton, 2000, Machine Design, Prentice Hall, Upper Saddle River, NJ.

[31] R. Juvinall and K. Marshek, 1991, Fundamentals of Machine Component Design,
2
nd
Ed., John Wiley and Sons, Hoboken, NJ.

[32] A. Burr and J. Cheatham, 1995, Mechanical Analysis and Design, 2
nd
Ed.,
Prentice Hall, Englewood Cliffs, NJ.

[33] N. Dowling, 1998, Mechanical Behavior of Materials: Engineering Methods for
Deformation, Fracture, and Fatigue, 2
nd
Ed., Prentice Hall, Upper Saddle River,
NJ.

[34] J. Bannantine, J. Comer, and J. Handrock, 1990, Fundamentals of Metal Fatigue
Analysis, Prentice Hall, Englewood Cliffs, NJ.

[35] R. Heywood, 1962, Designing Against Fatigue, Chapman and Hall, London,
England.

[36] B. C. Hanley and T. J. Dolan, 1953, Surface finish, ASME Handbook: Metals
Engineering Design, Vol. 3, pp. 100-106, McGraw-Hill, New York, NY.

[37] H. Wiegand, 1940, Effect of surface treatment on fatigue strength, MAP
Translation 1772, BMW, Flugmtorenbau, Berlin.


211
[38] R. Z. Manteuffel, 1941, Fatigue endurance of power vehicle springs and
possibilities of influencing same, Deut. Kraftfahrforsch, Vol. 49, pp. 1-51.

[39] F. Zimmerli, 1944, Shot blasting and its effect on fatigue life, Surface
Treatments of Metals, American Society for Metals, pp. 261-278.

[40] Nicholson, 1967, Heat Treatment of Steels, Vol. 61, Dept. of Scientific and
Industrial Research, Wellington, New Zealand.

[41] G. E. Totten, 2007, Steel Heat Treatment Handbook, CRC Press, Boca Raton, FL.

[42] G. E. Totten and M. Howes, 1997, Steel Heat Treatment Handbook, pp. 251-292,
Marcell Dekker, New York, NY.

[43] D. Callister, 2003, Material Science and Engineering an Introduction, 6
th
Ed.,
John Wiley and Sons, Hoboken, NJ.

[44] R. Landgraf, 1968, Cyclic deformation and fatigue behavior of hardened steels,
TAM Report, No. 320, pp. 1-98, University of Illinois, Department of Theoretical
and Applied Mechanics, Urbana, Illinois.

[45] J. M. Barsom, 1971, Fatigue-crack propagation in steels of various yield
strengths, Trans. ASME, J. Eng. Ind., Ser. B, No. 4, pp. 1190-1196.

[46] A. Ariel and A. Mukherjee, 1979, Decarburization and fatigue, The Enigma of
the Eighties: Environment, Economics, Energy, Vol. 1, pp. 103-111.

[47] P. Maiya and D. Busch, 1975, Effect of surface roughness on low cycle fatigue
behavior of type 304 stainless steel, Metallurgical Transactions A, Vol. 6A, pp.
1761-1766.

[48] G. Deng, K. Nagamoto, Y. Nakano, and T. Nakanishi, 2009, Evaluation of the
effect of surface roughness on crack initiation life, ICF12, pp. 1-8, Natural
Resources Canada, Ottawa, Canada.

[49] ASTM Standard E83-02, 2004, "Standard practice for verification and
classification of extensometers," Annual Book of ASTM Standards, Vol. 03.01,
pp. 232-244, ASTM International, West Conshohocken, PA.

[50] ASTM Standard E8-04, 2004, "Standard test methods for tension testing of
metallic materials," Annual Book of ASTM Standards, Vol. 03.01, pp. 62-85,
ASTM International, West Conshohocken, PA.

You might also like