You are on page 1of 207

A MATERIAL BASED APPROACH TO CREATING WEAR RESISTANT

SURFACES FOR HOT FORGING

DISSERTATION

Presented in Partial Fulfillment of the Requirements for


the Degree Doctor of Philosophy in the Graduate
School of the Ohio State University

By

Sailesh Babu, M. S.

*****

The Ohio State University


2004

Dissertation Committee: Approved by


R. Shivpuri, Adviser
S. L. Semiatin _______________________________
J. R. Brevick Adviser
Graduate Program in Industrial and Systems Engineering
ABSTRACT

Tools and dies used in metal forming are characterized by extremely high

temperatures at the interface, high local pressures and large metal to metal sliding. These

harsh conditions result in accelerated wear of tooling. Lubrication of tools, done to

improve metal flow drastically quenches the surface layers of the tools and compounds the

tool failure problem. This phenomenon becomes a serious issue when parts forged at

complex and are expected to meet tight tolerances. Unpredictable and hence uncontrolled

wear and degradation of tooling result in poor part quality and premature tool failure that

result in high scrap, shop downtime, poor efficiency and high cost.

The objective of this dissertation is to develop a computer-based methodology for

analyzing the requirements hot forging tooling to resist wear and plastic deformation and

wear and predicting life cycle of forge tooling. Development of such is a system is

complicated by the fact that wear and degradation of tooling is influenced by not only the

die material used but also numerous process controls like lubricant, dilution ratio, forging

temperature, equipment used, tool geometries among others. Phenomenological models

available in the literature give us a good thumb rule to selecting materials but do not

provide a way to evaluate pits performance in field. Once a material is chosen, there are no

proven approaches to create surfaces out of these materials. Coating approaches like PVD

ii
and CVD cannot generate thick coatings necessary to withstand the conditions under hot

forging. Welding cannot generate complex surfaces without several secondary operations

like heat treating and machining. If careful procedures are not followed, welds crack and

seldom survive forging loads. There is a strong need for an approach to selectively,

reliably and precisely deposit material of choice reliably on an existing surface which

exhibit not only good tribological properties but also good adhesion to the substrate.

Dissertation outlines development of a new cyclic contact test design to recreate

intermittent tempering seen in hot forging. This test has been used to validate the use of

tempering parameters in modeling of in-service softening of tool steel surfaces. The

dissertation also outlines an industrial case study, conducted at a forging company, to

validate the wear model. This dissertation also outlines efforts at Ohio State University, to

deposit Nickel Aluminide on AISI H13 substrate, using Laser Engineered Net Shaping

(LENS). Dissertation reports results from an array of experiments conducted using LENS

750 machine, at various power levels, table speeds and hatch spacing. Results pertaining

to bond quality, surface finish, compositional gradients and hardness are provided. Also, a

thermal – based finite element numerical model that was used to simulate the LENS

process is presented, along with some demonstrated results.

iii
Dedicated to my parents, Rekha and Rahul

iv
ACKNOWLEDGMENTS

The success of this research has been due to the invaluable contributions of various

individuals. First and foremost, I would like to thank Prof. Shivpuri for his vision and

support. I am also thankful for his guidance over the years, which have helped me see the

forest and not get lost amongst the trees.

I would also like to thank Prof. Brevick and Dr. Semiatin for serving on my

dissertation committee and providing valuable suggestions and guidance.

My heartfelt acknowledgements go to the following individuals and organizations,

for their help that has been instrumental in me finishing this doctoral work

• Marion Forge (Larry Shump, David Kuhlmann, Steve Grimes and others) for

giving me the first break in forging industry and for supporting me with tools, time

and tuition over a period of several years

• Dr. Tang and Dr. Wu at SFTC for granting me license for use of DEFORM 2D

and 3D for several years

• Staff at SFTC (John, Dave, Pavan, Chris and Jeff) for their support and advice

v
• Jim Seavers and John Canfield at Eaton Corporation, South Bend, Indiana for

providing me worn forging dies and process data for validation of models

developed

• Mary Hartzler and her IWSE machine shop staff, for their help in fabricating tools,

several times at short notice

• Lloyd Barnhart (Material testing Lab) and LENS lab staff (Dr. Pete Collins and

others) of Material Science department for their help in running test campaigns

• Finkl, Uddeholm, Advanced Heat treat, Balzers and Wright Patterson AFB for

their material donations

• All my colleagues, especially Dr. Satish Kini, Dr. Dilmar Ribeiro and Yuanjie Wu

for their assistance

I am thankful for the continued support and encouragement I have received from

my parents and my extended family. Lastly, I would like to thank my wife Rekha for the

understanding, support and motivation she has provided over the last several years, that

has helped me focus on finishing this work.

vi
VITA

February 8, 1969 Born – Madras, India

April, 1990 B. Tech, Mechanical Engineering, Indian


Institute of Technology, Madras.

December, 1992 M.S., Industrial and Systems


Engineering, University of Florida

March 1993 – May 1995 Graduate Research Associate, Industrial


and Systems Engineering, The Ohio
State University.

May 1995 – December 1999 Engineer, Eaton Corporation, Forge


Division

December 1999 – November 2001 Engineering Manager, Dana


Corporation, Forge Division

November 2001 – January 2003 Senior Manufacturing Engineer, Velocys


– A Subsidiary of Battelle Laboratory

January 2003 – present Research Engineer, The Ohio State


University

PUBLICATIONS

1. Shivpuri, R., Babu, S., Kini, S., Pauskar, P. and Deshpande, A., “Recent Advances in
Cold and Warm Forging Process Modeling Techniques: Selected Examples”, Journal of
Materials Processing Technology, Vol. 46, No. 1-2, pp. 253-274, 1994.

vii
2. Babu, S., Tan, B., Osborne, T. and Yeralan, S., “Product Quality/Tool Reliability
Modeling of Complex Manufacturing Tools”, Turkish Journal of Industrial Engineering,
Vol. 5, No. 3, pp. 3-11, 1995.

3. Babu, S., Tan, B. and Yeralan, S., “Computer Aided Reliability Modeling And
Applications In Semiconductor Manufacturing”, Computers & Industrial Engineering,
Vol. 23, No. 1-4, pp. 169-172, 1992.

FIELDS OF STUDY

Major Field: Industrial and Systems Engineering


Specialization in Manufacturing Processes

Minor Fields: Mechanical Design and Physical Metallurgy

viii
TABLE OF CONTENTS

Page

ABSTRACT.................................................................................................................... ii

VITA ............................................................................................................................ vii

LIST OF FIGURES ..................................................................................................... xiii

LIST OF TABLES ....................................................................................................... xxi

Chapters:

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

1.1 Problem Statement................................................................................................. 1

1.2 Research Objective ................................................................................................ 6

1.3 Research Approach................................................................................................ 7

1.4 Research significance and system benefits .............................................................. 8

1.5 Dissertation outline................................................................................................ 9

FORGING DIE FAILURE: A FUNDAMENTAL OVERVIEW.................................... 12

2.1 Failure in Metal Forming – Selected Examples ..................................................... 12

2.2 Various Modes of Tool Failure ............................................................................ 16

2.2.1 Plastic Deformation....................................................................................... 18

2.2.2 Abrasive wear ............................................................................................... 20

2.2.3 Thermal Fatigue ............................................................................................ 23

ix
2.2.4. Mechanical Fatigue ...................................................................................... 25

2.3 Die Life Improvement Strategies.......................................................................... 28

2.3.1 Die Materials ................................................................................................ 29

2.3.2 Heat Treatment ............................................................................................. 36

2.3.3 Surface Engineering ...................................................................................... 41

2.3.4 Hard coatings and advanced materials ........................................................... 55

2.4 Conclusions ......................................................................................................... 60

MODELING OF TEMPERING AND WEAR............................................................... 66

3.1 Overview of Past Studies in Modeling of Tempering, Wear and Die Life.............. 66

3.2. Characterization of Thermal Softening ................................................................ 76

3.3. Prior work in Wear Characterization................................................................... 78

FORMULATION FOR PREDICTING WEAR AND THERMAL SOFTENING .......... 87

4.1. Proposed Formulation for Thermal Softening and Wear of Hot Forging Tools .... 87

4.1.1 Intermittent Tempering Model Formulation................................................... 88

4.1.2 Architecture and Model Development ........................................................... 90

4.1.3. Formulation of Wear Model ......................................................................... 90

4.2. Demonstration of Thermal Softening Model........................................................ 93

4.2.1. Effect of lubricant heat ................................................................................. 94

4.2.2. Effect of Equipment ..................................................................................... 99

x
4.3 Summary ........................................................................................................... 101

VALIDATION OF WEAR AND OVER-TEMPERING.............................................. 102

5.1 Introduction....................................................................................................... 102

5.2 Thermal Softening Setup ................................................................................... 102

5.3 Thermal Softening Test Campaign ..................................................................... 109

5.4. Finite Element Modeling................................................................................... 113

5.5 Characterization................................................................................................. 117

5.6 Industrial Case Study......................................................................................... 119

INTEGRATED SYSTEM: FRAMEWORK AND APPLICATIONS........................... 131

6.1 Summary ........................................................................................................... 131

6.2 Future work....................................................................................................... 133

6.3 Conclusions ....................................................................................................... 134

DEVELOPMENT OF LENS PROCESS MODELS .................................................... 136

A.1 Introduction TO LENS ..................................................................................... 136

A.2 Prior Modeling Efforts ...................................................................................... 137

A.2.1 Stage 1: Powder free-fall in laser zone........................................................ 138

A.2.1. Stage 2: Melting of Substrate and bead formation...................................... 144

A.2.3. Stage 3: Mixing of Substrate and Powder .................................................. 150

A.2.4. Stage 4: Solidification ............................................................................... 152

xi
A.3 Numerical Model Development ......................................................................... 152

A.4 Parametric Numerical Study of LENS Process .................................................. 155

A.5 Validation - Experimental Procedure................................................................. 162

A.6 Characterization and Results ............................................................................. 166

A.6.1. Effect on surface finish .............................................................................. 166

A.6.2. Coating Morphology ................................................................................. 170

REFERENCES ........................................................................................................... 178

xii
LIST OF FIGURES

Figure Page

1.1. Typical precision forged gears with net teeth............................................................. 3

1.2. Approximate breakdown of costs due to tooling failure [Cser, et al., 1993]............... 4

1.3. Overview of iterative design approach ...................................................................... 8

1.4. Schematics of approach proposed in the dissertation ................................................. 9

2.1 Complex interaction of forging parameters and wear [Cser, et al., 1993] .................. 13

2.2. Example of abrasive wear in conventional closed die press forging – top blocker

punch, Finkl FX-2 dies after 2000 forging blows (courtesy - Dana Corp., Forge

Division, Marion, Ohio) ......................................................................................... 14

2.3. Example of wear and thermal fatigue (or heat checking) in upsetter inserts. Material is

FX-2 (courtesy - Dana Corp., Forge Division, Marion, Ohio)................................. 14

2.4. Wear marks typically found in forging of bevel gear dies (courtesy - Dana Corp.,

Forge Division, Marion, Ohio) ............................................................................... 15

2.5. Wear marks typically found in forging of bevel gear dies (courtesy - Dana Corp.,

Forge Division, Marion, Ohio) ............................................................................... 15

xiii
2.6. Flattened root and wear grooves in the direction of material flow in a H-13 spur gear

die (courtesy - Dana Corp., Forge Division, Marion, Ohio) .................................... 16

2.7. Frequency and location of typical die failures in forging (Cser et al. 1993)............... 17

2.8. Examples of hot forging die surfaces plastically deformed [Summerville, et al., 1995]

.............................................................................................................................. 19

2.9. Examples of surface plastic deformation [Summerville, et al., 1995]........................ 20

2.10. Hot forging top blocker punch made form H13 [Summerville, et al., 1995] ........... 21

2.11. S-N curve with probability lines or S-N-P [Dieter, 1986] ...................................... 26

2.12. Illustration of the methods for estimating fatigue based in static properties [Manson,

1972]..................................................................................................................... 28

2.13. Some aspects of forging and process design that affect wear and fracture [Cser, et

al., 1993] ............................................................................................................... 29

2.14. Heat treatment cycle of hot working steels [Krauss, 1995] .................................... 38

2.15. Thickness of various coatings and surface treatments [Subramanian, et al., 1996].. 42

2.16. Comparison of wear amounts of surface treated upsetting tools after 1000 forging

cycles with lubricant (Deltaforge-31) [Doege, et al., 1996]..................................... 42

2.17. Results for hot work tool steels in the H13 group presented by Krishnadev

[Krishnadev and Jain, 1997] (a) Composition (b) toughness (c) hot hardness (d)

softening of the alloys 2-3 with and without nitriding e) Hardness achievable with

xiv
different coatings and the alloys chemical composition f) Charpy impact toughness of

H-13 and treated alloy No. 3.................................................................................. 49

2.18. Comparison of yield strength of IC-15 to those of other high temperature alloys. .. 60

3.1. Various process and material .................................................................................. 67

3.2. Variation of die temperatures during a hot forging cycle ......................................... 68

3.3. Effect of forging temperature on the wear depth after forging 4000 pieces [Netthofel,

1965]..................................................................................................................... 70

3.4. Variation of wear volume with die bulk temperature for lubricated and dry forging

[Singh, et al., 1973] ............................................................................................... 71

3.5. Drop in toughness of H-13 with secondary hardening. a) represents toughness in

Charpy V-notch energy, b) represents toughness in Kic [Pickering, 1987] .............. 74

3.6. Master tempering curve for premium H13 were: P= Larson-Miller parameter, T is

temperature (°F), t is time in seconds [Carpenter, 1999]......................................... 79

3.7 Typical cylinder tests used to characterize wear and resulting measurable wear pattern

[Singh, et al., 1973] ............................................................................................... 80

3.8. Experimental setup used by Tittagala and others [Tittagala, et al., 1982]................. 81

3.9. Punch designs used by Bariani [Bariani, et al., 1996]............................................... 83

3.10. Dies used by Silva and Dean for wear characterization [Silva and Dean, 1971]...... 83

xv
3.11 Extrusion – type tests conducted by Doege, Melching and Kowallick [Doege, 1978].

.............................................................................................................................. 84

3.12. Doege’s experimental setup [Doege, 1994]........................................................... 86

3.13. Test setup used by Netthofel [Netthofel, 1965] ..................................................... 86

4.1. Schematics of modeling in-process hardness loss of forging tools............................ 89

4.2. 2 stage gear blank forging sequence (Courtesy: Sypris Technologies) ..................... 95

4.3. Hardness distribution after 3000 and 5000 shots, heat transfer coefficient used = 12

KW/m2°C, press type: mechanical press ................................................................. 96

4.4. Hardness distribution after 3000 and 5000 shots, heat transfer coefficient used = 24

KW/m2°C, press type: mechanical press ................................................................. 96

4.5. Hardness loss of the top die nose and flash areas after 3000 and 5000 strokes,

assuming a heat transfer coefficient of 33 KW/m2°C............................................... 97

4.6. Hardness distribution at the surface, after 3000 and 5000 shots, heat transfer

coefficient used = 12 KW/m2°C, press type: mechanical press ................................ 98

4.7. Hardness distribution after 3000 and 5000 blows, with a slow hydraulic press (p2)

v=10 mm/sec, h = 6 KW/m2 °C, m=0.3................................................................. 100

4.8. Hardness loss after (a) 3000 and (b) 5000 blows, with forging on a hydraulic press

with ram speed v=20 mm/sec, h = 6 KW/m2 °C, m = 0.3 ...................................... 101

5.1 Typical thermal cycle in a hydraulic and hammer forging........................................ 104

xvi
5.2. Temperatures spike seen at the die surface, in a typical hot forging die. Simulation

was performed in DEFORM2D............................................................................ 105

5.3. Schematics of the test setup to recreate the thermal cycle...................................... 106

5.4. a) Waspaloy billet mounted to press top bed b) assembled instrumented test container

mounted to the press ram (bottom bed)................................................................ 107

5.5. Eight H13 die samples assembled in a container .................................................... 108

5.6. Test setup on Instron 1322 ................................................................................... 109

5.7. Velocity profile used in the tempering test trials, positive indicates motion of ram

towards billet and negative is away from billet...................................................... 110

5.8. Thermal history of 0.010” below surface during contact test campaign.................. 112

5.9. Recorded temperature at .010” below the surface (10 cycles)................................ 113

5.10. Predicted hardness loss at the surface after 800 cycles......................................... 116

5.11. Predicted hardness distribution at the surface, after 1200 pieces .......................... 116

5.12. Predicted hardness distribution at the surface, after 1600 pieces .......................... 117

5.13. Measured micro-hardness measurements from surface to the interior .................. 118

5.14. Measured value of hardness loss versus predicted hardness loss at different depths

from the contact surface....................................................................................... 119

5.15. Schematics of a) buster, b) blocker c) finisher d) end of stroke and e)gear blank

forging (courtesy: Eaton Corp., South bend, Indiana)........................................... 121

xvii
5.16. (a) Worn blocker and finish dies (b) Close-up of blocker used in validation ......... 122

5.17. Predicted total wear profile in blocker die for the gear blank forging case study .. 124

5.18. Predicted total wear distribution from center to outside diameter, in blocker die.. 125

5.19. Corner wear predicted using modified Archard model......................................... 126

5.20. Worn die and Sheffield Cordax contact type CMM used to measure the profile of

the worn die......................................................................................................... 127

5.21. CAD approach used to compare the predicted and measured wear profile ........... 128

A.1. Microstructure v/s processing conditions for laser cladding (D = beam diameter in

mm, V is the table speed in mm/s)[Kathuria, 1997] .............................................. 143

A.2 Effect of travel speed on aspect ratio of the clad bead [Qian, et al., 1997] ............. 147

A.3. Variation of surface finish in powder-based cladding[Li and Ma, 1997] ................ 148

A.4. Effects of laser traverse speed on aspect ratio of bead and dilution [Ming, et al.,

1998]................................................................................................................... 150

A.5. Laser beam representation and thermal boundary condition modeled in FEM ....... 154

A.6. Typical thermal profile obtained from the simulation, used to obtain the length, width

and height of the melt pool................................................................................... 158

A.6. Typical thermal profile obtained from DEFORM-3D simulations of LENS process

(temperature in °C) .............................................................................................. 159

xviii
A.7. Simulated effect of absorbed power on weld superheat (table speed = 12.5 mm / sec,

preheat temperature = 20 °C, H13 substrate, laser beam diameter = 1 mm) .......... 160

A.8. Simulated effect of table speed on melt pool superheat (effective power = 120 watts,

preheat temperature = 20 °C, H13 substrate, laser beam diameter = 1 mm) .......... 161

A.9. Simulated effect of beam diameter on melt pool superheat (effective absorbed

power=120 watts, table speed = 12.5 mm/sec, preheat temperature =20 °C, H13

substrate)............................................................................................................. 161

A.10. LENS system a) schematic and b. experimental facility at the Ohio State University

[Banerjee, et al., 2002]......................................................................................... 164

A.11. Federal Surface Analyzer used in the surface measurement................................. 167

A.12. Profile – sample 2 (in microns) ........................................................................... 168

A.13. Profile – sample 4 (in microns) ........................................................................... 168

A.14. Profile – sample 5 (in microns) ........................................................................... 168

A.15. Profile – sample 7 (in microns) ........................................................................... 169

A.16. Profile – sample 8 (in microns) ........................................................................... 169

A.17. Profile – sample 9 (in microns) ........................................................................... 169

A.18. Optical images of coated surface at 50X for the samples 2-9. Note: samples 3, 5, 7,

and 9 were coated at higher power levels, samples 6,7,8 and 9 have .015 HS ....... 171

xix
A.19. SEM image showing cross section of sample 7, polished and bakelite mounted.

Sample was not etched a) 75X b)300X ................................................................ 173

A.20. SEM image of surface of NiAl coating on H-13 substrate showing random

appearance of partly melted powder. Image was taken at 75X and 381X.............. 173

A.21. SEM Image of sample 7, showing NiAl powder at different stages of melting.

Images are at a)150X b)1000X c) 1200X and b) 5000X....................................... 174

A.22. EDS analysis of a) Surface of sample 7 and b) NiAl powder ............................... 175

A.23. Variation of nickel, iron and chromium across the interface ................................ 177

xx
LIST OF TABLES

Table Page

1.1. AGMA Classification of gear quality for a typical automotive gear............................ 5

2.1. Class 510. Chromium die steel [Roberts, et al., 1998] ............................................. 31

2.2. Class 520. Chromium – Molybdenum die steels [Roberts, et al., 1998].................... 31

2.3. Class 530. Chromium – Tungsten die steels [Roberts, et al., 1998].......................... 32

2.4. Class 540. Tungsten die steels [Roberts, et al., 1998].............................................. 32

2.5 Composition of common maraging steels, VascoMax is a trade name of Teledyne, (*)

indicates maximum allowed content, (+) trademark of Crucible steel...................... 34

2.6. Hardening and tempering temperatures and procedures for tool steels [Roberts, et al.,

1998]..................................................................................................................... 39

2.7. Response of different tool steels to several surface engineering towards enhancement

of toughness, hot hardness, heat checking, temper resistance [Krishnadev and Jain,

1997]..................................................................................................................... 48

2.8. Average maximum wear depths (µm) on surface engineered dies after upsetting 500

AISI 1040 steel billets at 1070° C [Venkatesan, et al., 1998].................................. 52

4.1. Various wear models reported in the literature ........................................................ 92

xxi
5.1. Nominal composition of Waspaloy........................................................................ 107

5.2. Various heat transfer coefficients used in the model and corresponding maximum

temperatures at a depth of .010” from surface ...................................................... 114

5.3. Measured hardness loss at different depths from the surface. Measurement was made

using a microhardness tester under a load of 500 gms .......................................... 115

5.4. Hot hardness of H13............................................................................................. 123

5.5. Predicted, normalized and measured values of the wear in blocker die................... 129

A.1. Simulation matrix used for modeling LENS process in DEFORM-3D .................. 156

A.2. Simulated maximum melt pool superheat temperatures and melt pool sizes........... 157

A.3. Measured surface roughness and profile of LENS coated NiAl samples, powder size

used was 100-270 ASTM mesh............................................................................ 166

A.4. Measured surface roughness and profile of LENS coated NiAl samples................ 167

A.5. Compositional gradient (wt %) from surface to the bulk of substrate, obtained

through Energy Dispersive Spectrometric (EDS) analysis..................................... 176

xxii
CHAPTER 1

INTRODUCTION

1.1 Problem Statement

Closed die forging at high temperatures is one of the oldest known manufacturing

methods that have the capability of producing near net shape parts. The first known closed

die forgings made using gravity-assisted hammers in early 1800s had liberal stock

allowances and large tolerances. At that time, hot forging process was a tremendous

improvement over machining parts from bar stock or castings. Towards the middle of 20th

century, the process was extended to forge aerospace materials, stainless steel, nickel and

titanium based alloys that have much lower formability and higher flow strength. In

addition, the shape of parts forged became increasingly complex. Slowly but steadily, the

demands placed on the die and the die material increased. With the eighties and nineties,

because of intense competition from other forge shops as well as from other competing

processes like die-casting and cold forging, hot forgers have started making forgings that

have lower stock allowances, better surface finish and tighter tolerances. Tolerances of +/-

.030” and scaly finishes have been replaced by scale-free surface requirements, low

1
machining stocks and tight tolerances of up to +/- .005”. Precision gear forgings and

flashless forgings typify this trend.

Precision near-net shape manufacturing processes are processes that produce

product shapes close to final desired shape. Flashless forgings with less than .030”

machining stock, spur gears, spiral bevel gears and splines with near net tooth with as little

as .005” grinding stock and net toothed bevel gears have become part of many forger’s

product line. Figure 1.1 shows some typical precision forged gears.

With such small machining allowances and tolerances, there is very little room for

error arising due to forging process variations, design and tooling failure. The forgers have

to reduce the variations in billet volumes, forging temperature, lubricant spray and

equipment characteristics and reduce die wear as much as possible, so as to make parts

that consistently meet or exceed the customer’s requirements economically. The

importance of process control has been well understood by aerospace material forgers. Of

late, this awareness of process - quality interaction is creeping into automotive steel

forgers too.

2
Figure 1.1. Typical precision forged gears with net teeth

Quality and process capability aside, cost is an important factor that impacts the

viability of a product line. One of the most important ingredients in the cost of forgings is

the cost of tooling. Die costs range from 10 to 15% of the cost of a forging [Doege, et al.,

1996]. This includes cost of die material, machining the dies and subsequent heat

treatment, if necessary. The indirect cost of dies is however, far more significant. If tooling

wears out or become unusable, the production has to be stopped to change dies. Setup

times can range anywhere from under 10 minutes to over 3-4 hours, depending on the

complexity of the setup, skill and practices used by the setup crew. This results in

additional direct wages in material handling, tool rework and other overhead costs. Also,

this may result in additional overtime premiums in the die shop and the forge shop, low

3
resource utilization and in some cases, result in missed delivery to customers. If quality

and inspection systems used to monitor part quality, breaks down and if dies are not

changed at the appropriate time, additional losses occur due to scrap. The effect of tooling

failure on setup costs is shown in (Figure 1.2). Though tooling cost is only 10 - 15% of a

forging cost, the indirect cost of tooling could be as high as 70% in short production runs.

The life of a forge tooling, hence, has great ramifications on the economic competitiveness

of a forging company. Identifying different modes of die failure and understanding

dominant mechanisms are essential first steps in the path to increasing die life.

Figure 1.2. Approximate breakdown of costs due to tooling failure [Cser, et al., 1993]

4
One of the part families where cost and quality of parts play a key role in the

viability of the process is precision forged gears. Typical run-out, profile errors, lead

errors and index errors in different AGMA class gears are shown in Table 1.1. In this

table, index error is the displacement of any tooth from its theoretical position, relative to

a datum tooth. Lead error or tooth alignment variation, is the difference between the

measured tooth alignment and the specified tooth alignment measured normal to the

specified tooth alignment and the tooth surface on the functional face width. Profile error

is the permissible amount of profile variation in the functional profile compared to the

datum. The run-out error is basically the same as concentricity of the I.D. to the pitch

diameter of the gear.

Table 1.1. AGMA Classification of gear quality for a typical automotive gear

Due to tool wear, process variations and general tolerances achievable in hot

forging, gears of class 8 and higher cannot be forged. The difficulty in forging higher class

gears (6 and higher) is well documented in the literature [Dean, 2000, Doege and Nagele,

5
1994, Douglas and Kuhlmann, 2000]. Class 6 and 7 gears can be manufactured by forging

followed by cold coining or ironing. This operation irons out low levels of imperfections

in the teeth profile caused by scaling, tool wear and handling. If the profile errors of gears

immediately after the forging operation are too high, secondary cold coining operation

cannot effectively improve the gear class. High degree of process control and tool

engineering are required in forging of these fine tooth gears to make this possible.

Current state of knowledge and technology makes forging of higher class fine

tooth gears economically and technically challenging. Understanding the die life problems

relevant to precision hot forging of net gears and help overcome premature die failure

through innovative solutions are keys to opening new forging markets in this area. This

dissertation provides a brief background into precision forged gears and different modes of

tool failure typical in hot precision forging. Possible recipes for reducing die failure are

discussed. Selected die life characterization approaches currently available along with their

pros and cons are discussed. A model-based approach to assess tool life in hot forging

process is described in detail. A framework for computer program is outlined. Results

from experiments to validate the model are provided along with selected case studies to

demonstrate its use.

1.2 Research Objective

The objective of this work is to come up with a scientific method to design and

build hot forging dies for extended die life. The integrated design approach proposed

6
combines a numerical model of forging process to predict failure of dies through wear and

thermal softening, with advanced die materials to assist in building forging dies that can

provide superior performance to conventional die designs.

1.3 Research Approach

The key elements of the approach proposed and tested in this work, are as follows:

1. Development of a numerical model of material over-tempering: This element builds

a numerical model of tool life taking into account the intermittent tempering of

tools. Hardness is the predominant material factor that affects tool life in hot

forging.

2. Development of numerical model of wear: Conventional models, as discussed

later, does not take into account the effect of process on hardness nor do they take

into account the slow degradation of material during a forging run. This element

models behavior of material that can later be used in a modified Archard’s wear

model to predict and model hot forging die life.

3. Calibration of developed models for application in hot forging: Numerical models

developed above require calibration. This element of the dissertation calibrates the

softening behavior and lays the ground work for industrial validation.

4. Validation of model for an industrial forging application: Validation of the tool life

model developed through industrial tools.

7
Figure 1.3. Overview of iterative design approach

1.4 Research significance and system benefits

The research work presented here, provides a method to correlate die wear, tool

failure and hence die life to the forging process conditions. Because this is done through

the use of numerical models, forgers may be able to come up with new processes or

redesign existing processes and tools, to reduce tool wear, improve quality and hence,

reduce part costs. This approach also allows forgers to access and compensate for the

severity of forging processes for new products like fine tooth net and near net spur, bevel

and spiral bevel gears that are AGMA Class 7 or higher.

8
Other Inputs Failure Criterion
Temperature
Dimension
Lubrication
Targets
Equipment
Cost Targets
Work piece
Die Life
Die Material

Tool Design Life


Geometry, Assessment
Compute
Coating type FEM Modeling Acceptable?
Wear life
Location
Toughness
Thickness

Die Material
Behavior

Changes to
Coating Material,
Thickness,
Location

Figure 1.4. Schematics of approach proposed in the dissertation

This will allow the design engineer to select and specify appropriate die material

for applications based on the process. Die manufacturing and repair also benefit a great

deal from novel repair approach that is proposed and tested in this work. By carefully

selecting process parameters that are used to repair the dies, one can reliably repair dies so

that properties obtained are acceptable.

1.5 Dissertation outline

Chapter 1 outlines the motivation behind this work along with the objectives and

research approach.

9
Chapter 2 presents technical hurdles along with an overview of die materials,

coatings and surface treatments of steels. Principles of tool failure through wear, thermal

fatigue and plastic deformation are presented.

Chapter 3 discusses the microstructural changes to the tools during the forging

process due to tempering and how it impacts wear. This chapter will also present various

research efforts reported in characterizing wear and life of hot forging tools.

Chapter 4 will go over the formulation of the numerical model, developed for

analyzing tool life during hot forging. Demonstration of the generated code is also

presented. Chapter 4 also presents some finite element results that demonstrate the effects

of process variables on tool wear.

Chapter 5 presents experimental work performed to calibrate the tempering and

wear model. This chapter also presents results of validation performed on an industrial hot

forging.

Chapter 6 summarizes the conclusions of this work and discusses the scope for

future work.

The appendix describes one of the promising techniques for tool and die repair:

Laser Engineered Net Shaping (LENS). Advantages and disadvantages of the technique

are presented along with the underlying physics. The experimental and numerical models

developed for predicting weld parameters, the substrate hardening, weld dilution,

10
microstructure and mechanical properties of the clad layer. Appendix A also presents

results of deposition trials performed to validate and support the developed model.

11
CHAPTER 2

FORGING DIE FAILURE: A FUNDAMENTAL OVERVIEW

2.1 Failure in Metal Forming – Selected Examples

Several factors impact the physical underlying phenomena behind die failure -

pressures, sliding velocities, hot hardness and a variety of physical and mechanical

properties. From a process perspective, there are several variables that impact die wear

and tool failure. These include forging temperature, cycle times, use of protective

atmosphere during billet heating, lubricant spray and die preheating.

These factors interact in a very complex manner dictating the magnitude of wear

found in the forging process. For instance, increasing forging temperature reduces the

flow stress of the billet material but increases the heat transfer into the dies. While reduced

flow stress tends to reduce die pressure, the increased tool temperatures reduce the local

die hardness, reducing the wear resistance. Figure 2.1 illustrates some of the complex

interactions typically found in forging applications that impacts die failure. Several factors

like billet material grade, microstructure, equipment used and finish die geometry are not

controlled by the forger. However, judicious choice of lubricant application, die material

12
and coatings, forging temperature and other controllable factors could impact die life

positively. These choices impact the product cost immensely. Poor choices, in some cases,

like precision forged spur, bevel and spiral bevel gears of class 8 and higher, will make

forging an economically infeasible option.

Figure 2.1 Complex interaction of forging parameters and wear [Cser, et al., 1993]

13
Figure 2.2. Example of abrasive wear in conventional closed die press forging – top

blocker punch, Finkl FX-2 dies after 2000 forging blows (courtesy - Dana Corp., Forge

Division, Marion, Ohio)

Figure 2.3. Example of wear and thermal fatigue (or heat checking) in upsetter inserts.

Material is FX-2 (courtesy - Dana Corp., Forge Division, Marion, Ohio)

14
Wear

Figure 2.4. Wear marks typically found in forging of bevel gear dies (courtesy - Dana

Corp., Forge Division, Marion, Ohio)

Wear

Figure 2.5. Wear marks typically found in forging of bevel gear dies (courtesy - Dana

Corp., Forge Division, Marion, Ohio)

15
Flattening

Figure 2.6. Flattened root and wear grooves in the direction of material flow in a H-13

spur gear die (courtesy - Dana Corp., Forge Division, Marion, Ohio)

2.2 Various Modes of Tool Failure

It is clear that several modes of failure contribute to premature tool life in hot and

warm forgings. To increase tool life, there is a need to understand the different modes of

failure requiring a closer look at the underlying mechanisms. In hot forging, in general, the

main reasons for die failure are

• Abrasive, adhesive and oxidative wear of dies during production

• Mechanical fatigue and gross cracking of dies during production

• Thermal fatigue or heat checking

16
• Plastic deformation of corners and sharp edges

Figure 2.7. Frequency and location of typical die failures in forging (Cser et al. 1993)

Several of these could play a role in failure of forging dies. Figure 2.7 above,

illustrates the comparative severity of these mechanisms in hot forging. Obviously, wear is

the most dominant mechanism found in hot forging. Sharp corners that experience

extremely high pressures and temperature fluctuations during a forging cycle are features

that are most vulnerable to wear.

In several cases, two or more of these mechanisms act together to wear down the

die. In this chapter, various mechanisms of die failure are discussed briefly. Also, physical

and mechanical properties of die material that impact its resistance to the mechanisms of

17
failure are listed. These are essential to characterize and model the mode of die failure and

to find cures for its prevention.

2.2.1 Plastic Deformation

Plastic deformation of die surface occurs at regions of the die that are subjected to

extreme pressure and temperatures and experience long contact times. This occurs when

the local stresses result in die stresses exceeding the local hot yield strength of the die

material. Typical areas of the die that are prone to plastic deformation are sharp corners of

the dies and thin protuberances that trap a lot of heat during the forging process. 2.8 and

Figure 2.9 show typical surfaces that are subjected to plastic deformation. Since extreme

pressures and temperatures cause this mode of failure, increased local forging stresses will

increase the chance of plastic deformation. Consequently, all design and process criterion

that impact stresses and die temperatures have an effect on the plastic deformation of dies.

Of these, the forging temperature, size and geometry of the forging, lubricant used,

forging cycle times, type of equipment used and the type of forging (whether it is

conventional or flashless) are the most important factors. These parameters either increase

local stresses or reduce the strength of the die by thermal softening or a combination of

both.

Storen and others [Tulsyan, et al., 1993] provide a good criterion, given in

Equation 2.1, to avoid plastic deformation in forging dies. They say that one can avoid

deformation related die failure if we follow the following criterion.

18
σ Z p 0.75 × H B (2.1)

where HB is the Brinell hardness of the die material at the maximum temperature, σ Z is

the local normal pressures. In hot or warm forgings, the hardness levels changes over the

course of a run because of tempering effects. Also, the roughness of the surface, to some

extent, affects the heat transfer coefficient, heating and hence the hardness.

Figure 2.8. Examples of hot forging die surfaces plastically deformed [Summerville, et al.,

1995]

19
Figure 2.9. Examples of surface plastic deformation [Summerville, et al., 1995]

2.2.2 Abrasive wear

Abrasive wear arises when a hard, rough surface slides against a softer surface,

digs into it, and plows a series of grooves in the softer surfaces. The material originally in

the groves is removed in the form of loose fragments, or forms ridges along each groove.

The material in the ridges is then vulnerable to subsequent complete removal from the

surface [Stachowiak, 1993]. In hot forging conditions, abrasive wear is compounded by

the presence of hard particles in the interface (three-body wear). These particles may be

hard oxides or scales, external-contaminating particles or other hard carbides dislodged

from the die surface. Figure 2.10 illustrates abrasive wear grooves on the top blocker in

hot forging.

20
Abrasive wear results in the removal of die material from the surface. This abrasion

is typically, either caused by the presence of hard particles between the die and the

deforming billet or protuberances embedded in the billet. The hardness of the particle that

causes the initial groove has to be equal to or greater than the hardness of the die.

Abrasive wear is a die failure phenomenon where die surface looses its definition over a

period of time due to constant abrasive action of material sliding across the die features

under high pressures.

Figure 2.10. Hot forging top blocker punch made form H13 [Summerville, et al., 1995]

Sometimes the sliding metal tends to dissolve the surface features of the die. This

phenomenon is called adhesive wear. Adhesive wear in hot forging may start as localized

21
welding since the interface temperatures can be as high as 1200 °C. This phenomenon is

generally manifested in the die picking up portions of the billet material and is accelerated

when nascent metallic die surface comes into contact with the hot billet. This probably

occurs in the following sequence of events:

• The lubricant layers and oxidation layers in both the die surface and billet surface have

been removed by abrasive wear

• The base metal of the billet makes contact with the base die steel and

• Sliding reduces but the interfacial pressure increases

The billet and die material pair having chemical affinity for each other (for example,

aluminum and die steel) form welds or joints

The part is ejected from the die. Either a portion of the die material is removed

with the billet or a portion of the billet material adheres to the die. This second possibility

is more common since the die material is generally several times stronger than the billet

material. Oxidation of the die surface tends to enhance both the abrasive and adhesive

wear rates by increasing the surface exposed.

Several works exists in the published literature that tries to characterize and model

wear in hot forging. Others are based on process variables such as forging area, weights

and energy while others have taken a more fundamental approach to modeling. With

advances in finite element models and computing, it is possible to use fundamental

material properties and process variables computed from FEM software, to model wear

22
more comprehensively. With the technological capabilities and data available, it is possible

to use Archard’s model provided in equation 2.2 to model wear as a function of thermo-

mechanical history of dies during a forging process and the changing hardness of the die

material due to continuous tempering.

p i × Vi
wear = k ∫ dt
Hi (2.2)

where p is normal pressure at a die location, V is the sliding velocity at any time, H is

the hardness of the die location and k is a constant dependent on several factors like billet

material and scale formation.

2.2.3 Thermal Fatigue

The appearance of a fine network of cracks in the hot and warm forging dies is

known as heat checking. The hot and warm forging processes have a typical cycle that

causes heating and cooling of the dies surfaces. The billet at high temperature is

compressed into the die cavity causing a drastic increase in the surface temperature. The

temperature increase at the surface of the die causes its expansion. At the same time, the

lower temperature of the die block constrains the expansion, generating compressive

stress. Next, the part is ejected from the die and the dies are lubricated. During the

lubricant spray the surface is rapidly cooled and the expansion is reversed causing tensile

stress. When cracks are formed by repetitive change in temperatures the phenomenon is

called thermal fatigue.

23
Thermal fatigue is caused due to non-uniform temperature distribution between the

surface of dies and the interior. Any temperature differences between the surface and

interior results in strain differentials due to varying thermal expansion. When the resulting

stress at the surface exceeds the hot strength of the material we have yielding of surface

layers. Extended cycling will result in crack initiation and subsequent growth of thermal

cracks. If the maximum and minimum temperatures at a die location are known, then low

cycle fatigue life can be calculated as follows.


1−ν σ
 1 −ν σ
 

α T _ T  > 2  1  + 2  2


 2 1 E E
1 2 (2.3)

where α is the mean coefficient of thermal expansion, ν is the Poisson ratio, σ is the

stress, 1 and 2 indicate the maximum and minimum values of temperatures.

Crack initiation occurs when the following criterion (from Coffin-Manson,

equation 2.4) is met.

N n F ε p = Cε f
(2.4)

where N is the number of cycles to crack initiation, n is a material constant from 0 to 1,

ε p is the plastic strain range, C is a constant that is between 0 and 1, andε f is the true

deformation to fracture – a material property.

Crack growth occurs at a rate given by equation 2.5.

24
1−ν σ 1−ν σ
   
da = aρ ε q = aρ α T −T  −  1 1 −  2  2 ]
q
 p 
dN     2 1 E E
1 2 (2.5)

where a is the crack length, N is the number of cycles, ρ and q are positive constants

dependent on material.

Any physical or process factors that impact the strain difference, impacts heat

checking. It should be noted that in real life, all these different modes of failure occur -

one mode aiding or reducing the effects of another.

2.2.4. Mechanical Fatigue

Mechanical fatigue is caused by initiation and growth of cracks in the die due to

fluctuations in pressures and die stresses in certain regions of the die. Repeated loading

and unloading of dies result in the propagation of cracks. Figure 2.11 [Dieter, 1986]

illustrates a probabilistic S-N curve typically found in literature. S-N curves represents a

family of curves that can be used to predict the cycles to failure of a mechanical

component when the range of stress it is subjected to, is known.

25
Figure 2.11. S-N curve with probability lines or S-N-P [Dieter, 1986]

The predominant type of fatigue that is found in metal forming dies is low cycle

fatigue which is associated with high stresses and temperatures. Low cycle fatigue is

defined as mechanical fatigue failure that occurs after less than 1000 stress cycles. Low

cycle fatigue test results in general, are shown as plots of plastic strain range ∆εp against

number of cycles N. The plot of strain against number of cycles using a log scale for N

results in a straight line that is known as Coffin-Manson law. The relationship is shown in

equation 2.6.

The first model for strain controlled fatigue is known as Coffin-Manson law:

C −1 / 2
∆ε p = N f ∆ε p = = N f
−m
2 or (2.6)

26
where ∆ε p is the plastic strain, Nf is the number of cycles to failure, C and m are material

constants.

Manson later found a graphical method (universal slopes method) to evaluate

fatigue based on static tensile tests. The method uses relationship in equation 2.7 to model

the fatigue growth.

3.5σ u −1 / 2 − 0 .6
∆ε p = Nf + D 0 .6 N f
E (2.7)

where the first term is the elastic strain and the second term plastic strain. σ u is

conventional ultimate strength, E is the elastic modulus, Nf is the number of cycles to

failure and ε f (represented by D) is conventional logarithmic ductility

The graphical representation of this equation is shown in Figure 2.12, and it is a

very useful way to evaluate fatigue using static tensile test data.

27
Figure 2.12. Illustration of the methods for estimating fatigue based in static properties

[Manson, 1972]

2.3 Die Life Improvement Strategies

It is clear from the treatment above that die life improvement strategy should take

into account the process, the forged geometry and the predominant modes of failure.

Solutions to die life improvement revolve around selecting appropriate die materials,

surface treatments and coatings or clads. For instance, while H-13 and higher alloy tool

steels may help extend tool lives in hot forging operation performed in a mechanical press,

they are seldom used in hammer forgings. H-13 does not have the toughness to resist the

impact loads that are typical in hammer forging. The following section deals with various

strategies to improve die lives in hot and warm forging. Figure 2.13 shows the various

strategies that impact tool life in a concise form.

28
Figure 2.13. Some aspects of forging and process design that affect wear and fracture

[Cser, et al., 1993]

2.3.1 Die Materials

Die material selection is possibly the biggest factor that affects the life of dies in a

hot or warm forging operation. There is a large variety of tool steels available in the

market that can be used for hot and warm forging applications. These steels could be

categorized as low alloy tool steels (Groups 6G, 6F, 6H), air-hardening medium alloy tool

steels (A2, A7-A9), chromium hot work steel (H-10 – H-19), tungsten hot work steels

(H20-H26), and molybdenum hot work steels (H41-H43).

29
Selection of die material grade (steel composition and microstructure distribution)

and subsequent heat treatment play a key role in failure of dies. These properties

completely define the thermal and mechanical properties that affect the mode of failure and

the rate of tool failure. In this section, we will go over the main classifications of tool steel

grades and characteristics of incoming tool steel – alloying composition, physical and

mechanical properties. A short section will also discuss new non-steel based superalloys

available in the market.

A comprehensive classification of tool steels by the American Iron and Steel

Institute is provided by Roberts [Roberts, et al., 1998]. The groups are based on alloying

elements and applications. Steels that are not temperature resistant are generally not used

in making the forging dies. However, they used in other parts of the die set like the bolster

and spacers.

Hot work die steels are classified into 3 different categories [Roberts, et al., 1998]

based on their alloy content. These can be chromium based, tungsten or molybdenum

based or steels where tungsten and chromium are approximately in equal proportion. Most

hot work steels are low carbon steels with medium or high alloying elements.. Table 2.1

lists some of the common grades of chromium die steels. Table 2.2, 2.3 and 2.4 lists some

common grades of Chromium – Molybdenum, Chromium-Tungsten, Tungsten and

Molybdenum hot work steels [Roberts, et al., 1998] In modeling, analyzing and predicting

30
die failure, the knowledge of the physical and mechanical properties is very important.

Knowledge of these properties is necessary to both understand the reasons for die failure.

Table AISI C Mn Si W Cr V Mo
510 .95 0.30 0.30 4.0
511 .95 0.30 0.30 4.0 0.50 0.50
512 .60 0.30 0.30 4.0 0.75 0.50
513 S7 .50 0.70 0.30 3.25 1.40
514 .50 0.30 0.90 3.25 0.25 1.40

Table 2.1. Class 510. Chromium die steel [Roberts, et al., 1998]

Type AISI C Mn Si Cr Ni V Mo W
520 H-11 0.35 0.30 1.00 5.00 0.40 1.50
521 H-13 0.35 0.30 1.00 5.00 1.00 1.50
522 H-12 0.35 0.30 1.00 5.00 0.40 1.50 1.50
523 0.40 0.60 1.00 3.50 1.00 1.00 1.25
524 H-10 0.40 0.55 1.00 3.25 0.40 2.50
525 0.35 0.30 1.00 5.00 2.00

Table 2.2. Class 520. Chromium – Molybdenum die steels [Roberts, et al., 1998]

31
Type AISI C Mn Si Cr V W Mo Co
530 H14 0.40 0.30 1.00 5.00 0.25 5.00 0.25 0.50
531 H19 0.40 0.30 0.30 4.25 2.00 4.25 0.40 4.25
532 0.45 0.75 1.00 5.00 0.50 3.75 1.00 0.50
533 0.35 0.60 1.50 7.25 7.25
534 0.45 0.60 1.50 7.25 7.25
535 H16 0.55 0.60 0.90 7.00 7.00
536 H23 0.30 0.30 0.50 12.00 1.00 12.00

Table 2.3. Class 530. Chromium – Tungsten die steels [Roberts, et al., 1998]

Type AISI C Mn Si Cr Ni V Co W Mo
540 H21 0.35 0.30 0.30 3.50 0.50 9.00
541 H20 0.35 0.30 0.30 2.00 0.50 9.00
542 0.30 0.30 0.30 2.75 1.75 0.30 10.00 0.25
543 H22 0.35 0.30 0.30 2.00 0.40 11.00
544 0.30 0.30 0.30 2.50 0.40 3.60 12.00
545 H25 0.25 0.30 0.30 4.00 1.00 15.00
546 0.40 0.30 0.30 3.50 0.40 14.00
547 H24 0.45 0.30 0.30 3.00 0.50 15.00
548 0.35 0.30 0.30 4.00 2.50 14.00 2.00
549 H26 0.50 0.30 0.30 4.00 1.00 18.00
550 H15 0.35 0.30 0.40 3.75 0.75 1.00 6.00
551 H15 0.40 0.30 0.50 5.00 0.75 1.00 5.00
552 H43 0.55 0.30 0.30 4.00 2.00 8.00
553 H42 0.65 0.30 0.30 4.00 2.00 6.40 5.00
554 H41 0.65 0.30 0.30 4.00 1.00 1.50 8.00
555 0.30 0.50 0.30 3.00 3.00

Table 2.4. Class 540. Tungsten die steels [Roberts, et al., 1998]

32
Apart from these, because of practical reasons, they need to possess good

machinability and resistance to warping during heat treatment. Die material’s resistance to

plastic deformation depends on how well it retains its hardness with temperature. It also

depends on its yield strength. Resistance to mechanical shock relies on the material having

good fracture toughness commonly measured in Charpy V-notch testing units. Resistance

to wear depends on tempering resistance with temperature indicated by hardness

measurements at elevated temperatures. Resistance to heat checking depends on the

material having high ductility, good tempering resistance, high yield strength and low

thermal expansion. High heat conductivity and low thermal expansion coefficient in die

materials is desirable because it reduces the temperature gradient or associated thermal

strains which is the cause of thermal fatigue and shock. It is also desirable that the steel

retains all its properties for an extended period under elevated temperatures. The

resistance of a die steel to thermal softening mainly depends on its alloying constituents

and its distribution. The tempering characteristics of these tool steels obtained under

laboratory condition represents very well the die material’s resistance to thermal softening.

Maraging steels are relatively new group of steels that was primarily developed for

aerospace applications. It has high nickel, cobalt and molybdenum content but very little

carbon. After austenitization and quenching the steel, the structure is soft nickel martensite

or similar soft structure with typical hardness of 30 – 40 Rc. Aging this matrix at

temperatures around 500° C results in dispersed precipitation of intermetallic phases. This

precipitation is not concentrated at the grain boundary alone. This dramatically increases

33
the strength without unduly affecting the toughness. Some examples of maraging steels

commercially available are shown in Table 2.5.

Type Ni Co Mo Ti Al C* Si* Mn* S* P*


I- 18.5 8.5 3.25 0.2 .1 .03 .10 .10 .01 .01
II- 18.5 7.5 4.8 0.4 .1 .03 .10 .10 .01 .01
III- 18.5 9 4.8 .6 .1 .03 .10 .10 .01 .01
IV- 18 11.8 4.6 1.35 .1 .03 .10 .10 .01 .01
HWM 2 11 7.5 - - .05 .10 .10 .01 .01
X2NiCo 12 8 8 .5 .5 .03 .10 .10 .01 .01
Marlock( 18.0 11.0 5.0 0.3 0.01 0.1 0.01 0.01

Table 2.5 Composition of common maraging steels, VascoMax is a trade name of

Teledyne, (*) indicates maximum allowed content, (+) trademark of Crucible steel

Its high resistance to thermal shock and high toughness makes Maraging steels

good candidates for dies where the mode of failure is heat checking. Maraging steels, used

in die casting industry, are not very common in forging industry.

Nickel, cobalt and iron based superalloys are another group of die materials that

has excellent potential in hot precision forging. This group of materials has extremely high

temperature strength and thermal softening. Like maraging steels, this group of materials

34
gets its strength from precipitation strengthening of intermetallic compounds like Ni3Al.

There are 4 primary groups of superalloys. They are:

Iron-based alloys. This group comprises of die steels like H-46 and Inconel 706 and

contains over 12% of Chromium. Small amounts of Molybdenum and Tungsten provide

the matrix with high temperature strength. Iron based superalloys also include austenitic

steels with high chromium and nickel content. This group can be used in applications

where dies could heat up to 1200°F.

Nickel-Iron based alloys. This group of alloy contains 24-27% nickel, 10-15% chromium

and 50-60% iron along with small quantities of Molybdenum, Titanium and Vanadium.

The carbon content in these alloys is very small, typically less than .1%.

Nickel based alloys. This group of alloys contains virtually no iron. The primary

constituent of these alloys are nickel (50-80%), chromium (20%) and combination of

molybdenum, aluminum, tungsten, cobalt and columbium. These grades again, get their

strength from solid solution strengthening and can be put to service at temperatures up to

2200° F. Example of nickel-based superalloys are Waspalloy, Udimet 500 and Inconel

718.

Cobalt based alloys.This group of alloys are more ductile than the other groups. Again,

these are age hardenable alloys whose primary constituents are Nickel, Iron, Chromium,

Tungsten and Cobalt. These can be used in applications where it could reach 1900° F.

35
2.3.2 Heat Treatment

Once a die material is purchased to specification and the die cavity manufactured,

the heat treatment it is subjected to has a dramatic effect on the properties of the die

material. Understanding the effects of various elements of heat-treatment should guide the

heat-treat specification on the die.

Before selection of a die material for an application, it is imperative to know what

a material is capable of and the performance levels we can aspire for. Alloying elements

like Chromium, Nickel, Vanadium and Molybdenum play a dramatic role in determining

the range of physical and mechanical properties one can expect from the material. If

properly heat-treated, high alloy die materials are capable of delivering high wear

resistance by retaining its hardness at higher temperatures. In general, if a lower alloy die

material is used, for similar toughness, the wear resistance that one can expect would be

lower. Any attempt to heat treat the material to a higher hardness could potentially lead to

catastrophic failure of dies by reducing the toughness. On the other hand, if we do not

make use of the wear resistance a die material is capable of by heat-treating it to high

toughness values, we may not be utilizing the full potential of the material.

Heat treatment of die steels involves the following steps:

1. After dies are made, it is heated to austenitizing temperature. Austenitizing

temperatures for hot work tool steels range anywhere from 1000° C - 1500° C.

36
During this phase, the structure of steel transforms from ferrite-pearlite structure

to austenite.

2. The dies are held at these temperatures for an extended period. This is the “soak”

or “hold” time. During this stage, the structure becomes uniformly austenitic.

Carbides of alloying elements go into solution.

3. After soaking, the dies are quenched in a quench medium to temperatures below

the transformation temperature. During this phase, based on the cooling rate

different regions of die experiences, transforms into different phases. Martensite is

the ideal final structure, however in practice lower bainite, upper bainite, pearlite

or retained austenite can be present in the structures, specially in blocks with big

section.

4. Tempering is the next stage of heat treatment. Here, martensite formed as a result

of quenching is tempered to a tougher structure. This could be done in more than

one step to maximize the toughness achieved, without sacrificing hardness. These

stages in heat treating a tool steel die are illustrated in Figure 2.14. Table 2.6

shows the hardening and tempering temperatures and procedures for various tool

steels.

37
Figure 2.14. Heat treatment cycle of hot working steels [Krauss, 1995]

38
Continued

Table 2.6. Hardening and tempering temperatures and procedures for tool steels [Roberts,

et al., 1998]

39
Table 2.6. continued

40
2.3.3 Surface Engineering

Surface-engineering of dies are techniques and processes used to induce, modify

and enhance the surfaces properties giving it more wear, corrosion and fatigue resistance.

These techniques do not modify the soft and tough interior of dies. Figure 2.15 shows the

typical surface depths of various surface treatments. Die coatings and surface treatments,

used in forging industry, primarily increase the abrasive wear resistance of dies by

increasing the hardness of the surface layers of the die. Figure 2.16 shows some results

from forging experiments that clearly illustrate the efficacy of surface treatments. These

results were obtained from forging trials performed eccentric crank press. This section lists

different surface treatments applicable to precision forging applications and issues one

need to be aware of that may affect the die performance.

Most surface treatments used in dies and tools are diffusion–based. These

processes rely on diffusion of chemicals into the surface, modifying the surface chemistry

and the mechanical properties of the surface layer. The thickness of the surface treated

layer in these types of diffusion processes rely on the time and temperature at which the

hardening is performed. The time-temperature dependence is of the form shown in

equation 2.8.

D=K t (2.8)

where D is the depth of the hardened case, K is a temperature dependent constant and t is

the time of exposure.

41
Figure 2.15. Thickness of various coatings and surface treatments [Subramanian, et al.,

1996]

Figure 2.16. Comparison of wear amounts of surface treated upsetting tools after 1000

forging cycles with lubricant (Deltaforge-31) [Doege, et al., 1996]

42
Different diffusion based surface hardening techniques which may be applicable to

forging dies are

• Carburizing and pack cementation

• Nitriding

• Carbo-nitriding and Nitro-carburizing

• Boriding

• Toyota Diffusion Process

• Thermo-reactive diffusion

Carburizing is the process of adding carbon to low carbon steels. Not typically

used for forging dies, the process relies on heating the parts to high austenitizing

temperatures of over 1500 °F and exposing the surface to a carbon rich atmosphere.

Carbon diffuses into the austenitic surface of the parts, which are then quenched to

provide a martensitic structure on the surface. As discussed before, martensite has

excellent wear resistance. Coupled with the soft and tough core, this surface treatment

gives the parts good resistance to mechanical shock as well as wear.

Case depths and hardness levels achievable are dependent on the time of exposure

and the richness of carbon at the surface. Prolonged exposure to carbon-rich atmosphere

results in a deep case. However, the surface may have excessive retained austenite and

free carbides, which in turn will result in excessive residual stresses.

Based on the medium used, carburizing can be any of the following.

43
• Gas carburizing

• Vacuum carburizing

• Plasma carburizing

• Salt bath carburizing

• Pack carburizing

Of these, gas carburizing is the most commonly used process because of ease of

process control and low equipment costs. Pack carburizing, the uses a solid carburizing

pack, is also widely used. Carburizing has limited application if precision forging dies

because, dies materials used are medium to high alloy steels. The process has advantages

of flexibility and low cost for low production. However, labor cost for cleaning and

environmental restrictions make the gas or liquid process cheaper. The gas and plasma

carburizing process are also more controllable.

Nitriding, similar to carburizing, is a process of hardening the surface by diffusing

nitrogen into the surface. Nitriding processes are performed at temperatures between 925

and 1050 °F (500 to 550 °C) [Asm, 1964] where the structure is still ferritic. The process

results in formation of and outer case of Fe3N and a inner layer that is strengthened by a

solid solution of N. In some cases, a white layer of Fe4N is formed. This layer, also called

the “white layer”, may easily spall during use and has to be avoided.

Steels nitrided are typically medium carbon steels with strong nitride-forming

elements like aluminum, chromium, vanadium and molybdenum. It is important that

tempering of the die steel be performed at a temperature exceeding the nitriding

44
temperature prior to nitriding in order to optimize the property combination of the core

and the surface of the dies. Also, because of the low nitriding temperatures, there is

generally little distortion from this heat treating process.

Although the depth and hardness of the nitride case depends a great deal on the

nitriding time, these properties (particularly the hardness) are sharply dependent on the

composition of the steel as well. Die steels containing large amounts of strong nitride

formers such as chromium, vanadium, and molybdenum form shallow, very hard surface

layers. On the other hand, low-alloy chromium-containing die steels (such as 6G, 6F2)

form deeper surface layers which are tougher, but not as hard.

There are many techniques for nitriding: gas-nitriding, liquid-bath nitriding, ion-

nitriding, etc. Each of these will be discussed separately.

Gas Nitriding: Gas nitriding is a surface hardening process in which nitrogen is

introduced into the surface layers of ferrous materials by holding them in contact with a

nitrogen-containing gas, which is usually, ammonia. Because a brittle, nitrogen-rich layer

(the "white nitride layer") is produced by this process, the depth of the nitrided case is

usually kept small. Sometimes, a special two-stage gas-nitriding process, which minimizes

the depth of this layer, is employed [Weist and Westheide, 1986]

Liquid nitriding: Nitriding in a liquid salt bath, or liquid nitriding, is performed at

the same temperatures as gas nitriding, approximately 925 to 1050 °F (496 to 566 °C), but

typically requires less time than conventional single-stage gas nitriding. The salt baths

45
consist primarily of mixtures (in varying proportions) of sodium and potassium cyanide

(from which the nitrogen is released during nitriding) and sodium carbonate, potassium

carbonate, and potassium chloride. These baths result in cases containing both nitrogen

and carbon compounds. Modifications of this heat treating procedure include a process

involving aeration. [Asm, 1964] This leads to a less brittle case of Fe3N compared to gas

nitriding process which develops cases containing very brittle iron compounds richer in

nitrogen (e.g., Fe2N). Commercially, liquid bath nitriding processes such as Tufftriding

process have been used on metalworking tooling.

Ion-nitriding, also known as plasma nitriding and glow-discharge nitriding, is yet

another form and probably the most recently developed method of alloying the surface

layers of ferrous parts with nitrogen. [Taylor and Tookey, 1981], [Edenhofer, 1976]. In

this process, a part to be nitrided is placed in a reaction vessel into which an atmosphere

containing nitrogen and hydrogen are introduced. The part is electric resistance heated to

930 °F. It is made the cathode and the reaction vessel the anode in an electric circuit. A

glow discharge between the vessel and the part causes ionization of the gases, causing

nitrogen ions to impinge upon the surface of the part. Because these ions have much

greater energy than those in gas or liquid nitriding, penetration and thus surface nitriding is

much quicker in ion-nitriding. Among other advantages of this form of nitriding is the

ability to control and minimize the extent of brittle "white-layer" formation. In fact, with

proper atmosphere control, nitriding surfaces of Fe4N can be formed. This compound is

very ductile, and thus parts with hard, wear-resistant, and tough surfaces may be

46
produced. The major drawback of this method is the need for a special reaction vessel

whose size limits the size of parts that can be treated.

Nitriding can be used in conjunction with alloy selection to selectively enhance

resistance to certain modes of failure. In Table 2.7, Krishnadev [Krishnadev and Jain,

1997] presents efficiency of various surface treatments in increasing resistance to

mechanical failure, thermal softening, heat checking and wear.

Krishnadev [Krishnadev and Jain, 1997] also presents toughness, hot hardness and

softening characteristics of nitrided and non-nitrided of H-13 and alloys of similar

composition. Figures also show the different levels of hardness and toughness achievable

with different coatings and nitriding.

47
Table 2.7. Response of different tool steels to several surface engineering towards

enhancement of toughness, hot hardness, heat checking, temper resistance [Krishnadev

and Jain, 1997]

Efficacy of nitrided tools were tested by Dean and others [Dean and Sturgess,

1987] using extrusion type testing. Relative wear rates of nitrided and non-nitrided tool

steels in extrusion they obtained are shown in Figure 2.17. Dean indicates that nitriding

reduces wear as much as 50%.

48
(a)

(b)

(c) (d)

Continued

Figure 2.17. Results for hot work tool steels in the H13 group presented by Krishnadev

[Krishnadev and Jain, 1997] (a) Composition (b) toughness (c) hot hardness (d) softening

of the alloys 2-3 with and without nitriding e) Hardness achievable with different coatings

and the alloys chemical composition f) Charpy impact toughness of H-13 and treated alloy

No. 3

49
Figure 2.17. continued

(e)

(f)

Carbo-nitriding and nitro-carburizing are diffusion – based surface treatment

techniques that combine the effects of nitriding and carburizing. Carbo-nitriding relies on

hardening carburized austenitic surface layers using nitrogen as an agent. It is performed

at temperatures where austenite is stable. Nitrogen and carbon diffuses into the surface.

Carbon, on quenching, provides the superior hardness levels. Nitrogen increases the

hardenability of the surface, thus reducing the need for drastic cooling and subsequent

50
high distortion. It should be noted that nitrogen, like carbon is an austenite stabilizer.

Excessively rich carbo-nitriding medium and prolonged times may result in high nitrogen

concentrations at the surface that may result in high levels of retained austenite. Retained

austenite is detrimental to hardness and wear resistance.

Austenitic nitro-carburizing relies on formation of Fe3N (ε carbo-nitrides) at the

surface to improve the hardness levels. It is carried out at temperatures of 675 °C to 775

°C. Because the mechanism relies on the formation of carbo-nitrides and not carbon or

nitrogen trapped in surface matrix, there is no need for quenching. This process, hence,

results in low distortions.

Ferritic nitro-carburizing displays superior fatigue resistance and relies on both

formation of ε carbo-nitrides and diffusion of carbon and nitrogen into the substrate. The

process is carried out at temperatures where the surface is still ferritic – around 570 °C.

Diffused nitrogen is trapped in a solid solution by quenching in oil. The white layer of ε

carbo-nitrides improves the wear resistance while the diffused layer results in

improvements in fatigue strengths upto 120% [Asm, 1964]. The yield strength of the

substrate surface also increases. For this treatment to be effective, the surface layers need

to be clean of oxides, scales, oil or other contaminants. Vapor degreasing or grit blasting

with fine abrasives may be necessary steps to achieve the most out of ferritic nitro-

carburizing process. Various processes exist like Nitemper process, Alnat-N process and

51
black nitro-carburizing that rely on controlling the composition of white layer formed. For

details on this processes, refer to ASM handbook [Asm, 1964].

Venkatesan and others [Venkatesan, et al., 1998] have evaluated nitro-carburized

dies and compared its performance with quench and tempered H-13 and ASM 6F3,

Nitrided H-13, Borided and Vanadized (TD) H-13. Results obtained from their tests are

provided in Table 2.8. They found that nitro-carburized dies performed very well

compared to untreated H-13. Details of the process they used is unknown.

H13 Dies 6F3 Dies


Treatments Top Bottom Top Bottom
Q&T 46 110 156 236
Nitro-carburized 4 5 5 37
Nitrided 10 12 11 9
Borided 5 6 0 0
Vanadised 0 0 0 0

Table 2.8. Average maximum wear depths (µm) on surface engineered dies after upsetting

500 AISI 1040 steel billets at 1070° C [Venkatesan, et al., 1998]

Boriding or boronizing process relies on diffusion and subsequent absorption of

boron atoms in the metallic lattice and formation of interstitial boron compounds to harden

the structure. Diffusion treatment can be carried out in either a gas, molten salt, or pack

media at a temperature between 700 °C to 1000 °C, depending upon the process and the

52
material to be borided. Extremely hard-surface layers ranging from 11450 to 5000 HV

that has a low coefficient of friction are formed if the base metal forms borides. The

process does not require quenching. If the base material has to be heat treated, the heat

treatment can be done after boriding, although care is required to reduce quenching

stresses to prevent spalling of the borided layer. Borided layers resist thermal softening

better than nitrided layers. They also exhibit moderate to high resistance to oxidation.

However, boriding provides marginal increase in fatigue endurance limits.

Boriding of steels is also done electrolytically. Boron atoms are electro-deposited

onto the metal from a bath of molten salt containing fluorides of lithium, sodium,

potassium and boron. The dies are borided in the 1470 F (800 C) to 1650 F (900 C)

temperature range in an atmosphere of argon or a mixture of nitrogen and hydrogen.

Thickness of coating is from 0.0005 to 0.002 in (0.013 to 0.05 mm), and treatment lasts

15 minutes to 5 hours [Fiedler and Dobnar, 1972].

It has been stated that boriding results in undesirable interaction with alloying

elements of hot-work die steels (H series) and develops a soft layer [Burgreev and Dobar,

1972]. Porosity in the borided layer can develop for steels, which require post-boriding

heat treatment. For this reason, it is preferred to limit boriding to those alloys that do not

require further high-temperature treatment. For example, A6 (075 C, 20.0 Mn, 0.3 Si, 1.0

Cr, 1.35 Mo) air-hardening steel can be hardened from the boriding temperature by

cooling in air, and only requires tempering. This steel, therefore, can be safely borided.

53
Boriding process is also known to improve the wear resistance by forming borides

with the subsurface die steel. Shivpuri and Semiatin [Shivpuri and Semiatin, 1988] report

work by Vincze who borided dies at 900° C for 3 hours followed by quench and temper.

Vincze reported an increase in wear resistance of borided dies by 70% compared to

untreated dies. Burgreev and Dobnar [Burgreev and Dobar, 1972] also report large

increase in hammer die forging die life when boriding is used.

As reported earlier, Venkatesan and others [Venkatesan, et al., 1998] also report

enhanced wear resistance of borided tools compared to untreated dies. Boriding low alloy

steels result in a jagged boride layer that are deeper than boride layers formed in high alloy

steels. This is because, alloying elements reduce the diffusion of boron into the substrate.

This may explain results obtained by Venkatesan who found that performance levels of

borided H-13 are comparable to nitro-carburizing. However, they found that the efficacy

of boriding was higher with low alloy die steels like 6F3.

Thermo-reactive processes also called Toyota diffusion process (TD process) is

another diffusion type process that relies on forming hard carbides of V, Cr and Nb on the

surface of dies. The process is performed by placing preheated dies in a molten borax bath

at temperatures from 850 °C to 1050 °C for times ranging from ½ to 10 hours. The bath

also contains strong carbide forming elements like Niobium, Vanadium, Titanium, and

Chromium. Unlike other processes discussed earlier, TRD process results in buildup of

surface of carbides. Also, Vanadium and chromium diffuse into the steel substrate forming

iron-chromium and iron-vanadium solid solution layers beneath the carbide layer. After

54
treating, the die is quenched in air, salt or oil and tempered. Drastic quenching may cause

unacceptable distortion and needs to be avoided. To reduce distortion, it is preferable to

pre-machine and grind the dies before TRD processing. Best results are obtained for steels

with atleast .3% C.

Dies processed by this method have excellent wear resistance and resistance to

corrosion and oxidation. Arai [Arai, 1992, Arai, 1995, Arai, et al., 1995, Arai and Iwama,

1981, Arai and Komatsu, 1982, Tsuchiya, et al., 1997] has done a lot of work in validating

the use of TRD process in cold and hot forging. By coating 3” flashless dies made of SKD

62 with Cr Carbide using the TD process, he was able to double the life of dies. The life of

dies increased from 5000 to 10,600. Arai also reports that TD treated steels did not peel

or crack under repeated blows with a pointed hammer. Under similar conditions, TiC

coated by PVD or CVD cracked after 50,000 pieces and peeled after 1000,000 pieces.

Venkatesan and others [Venkatesan, et al., 1998] found that vandalized dies

showed the least wear of the tested specimen compared to nitro-carburizing and boriding.

They also noted that vanadized dies showed no traces of wear irrespective of the type of

substrate used. They also noted that nitro-carburizing and nitriding resulted in similar wear

rates for both types of die steels used in the study.

2.3.4 Hard coatings and advanced materials

The hardfacing is a coating process that applies a surface deposit that

metallurgically bonds to the base material. In the past, the process was used primarily for

55
repair and maintenance of dies and molds. Now, it is increasingly being used as an

inexpensive means for depositing a hard layer on localized wear-prone die areas. Nugent

[Nugent, 1986] reports in his study that weld deposits of Alloy 625 increased forging die

life by 400%.

Hardfacing or welding alloys are generally based on iron, cobalt and nickel metals.

Hard phases are formed by addition of carbon (in Fe) or boron (in Ni). The preferred

application methods for various alloys are: iron alloys deposited by surface weld methods,

cobalt alloys by welding and powder surfacing and hard nickel alloys in the form of

powder. The volume fraction for hard phase is very important for the wear resistance in

the weld deposit. Often there is no proportional dependence and the best wear resistance

is not achieved by the highest hard phase concentration.

Various ferrous alloys are used to repair steel dies or to lay down deposits of

better wear and heat resistance in the welded deposit. Often there is no proportional

dependence and the best wear resistance in not achieved by the highest hard phase

concentration. Different microstructural combinations are used to increase wear resistance

of tool steel, these include transformation behavior (bainite, eutectic) and the use of

carbide forming elements where chromium is used as alloying element. Austenitic and

austenitic-ferritic material is preferred for wear resistance under higher loads.

Hard-faced tool steels have to be heat treated before use. However, hyper-eutectic

cast or carbide sinter alloys are not suitable for heat treatment and the weldment from

56
carbide filler rods exhibit the required material properties directly after welding. In respect

of the economic importance, hard facing with iron base alloys predominates in comparison

with nickel and chromium alloys [Farmer, 1979]. This is more relevant with the increasing

automation of hard surfacing and the use of robots in welding systems.

Nickel- and cobalt-base alloys are the usual choices for hard-facing of dies.

Questions concerning the transformation or primary phase instability during hard surfacing

process can be considered of secondary importance in hard cobalt or nickel alloys. The

material properties are present after solidification from melt. Use of these alloys in

hardfacing offers a considerable saving over die blocks of these alloys. In a typical

hardfacing operation, one or two layer of alloy, each about 0.010 to 0.050 in. (0.25 +to

1.27 mm) thick are deposited in the die. If a very large amount of buildup is desired or

require, however, it is advisable to apply layers of stainless steel, high nickel alloy, or low-

alloy filler metal first rather than many layers of hardfacing material [Deevi and Sikka,

1996].

Hard nickel alloys are processed generally as metal powders (P/M) and to a lesser

extent as cast rod and electrode. Some nickel base alloys are applied as layers. With

several alloys, the hardening during loading is used to increase the wear strength, e.g. for

cladding cutting tools and die blocks for hot working.

The weld deposits used could be one of the following.

• Deposits of identical material onto a die block to repair it or to allow resinking

57
• Deposits of higher alloy steels (e.g., chromium hot-work steels) onto the die surface of

low-alloy steels to improve the service performance of the dies (e.g., wear and heat

resistance).

• Deposits of hard or high-temperature materials (usually cobalt or nickel-based alloys)

onto low-alloy or hot-work steels to improve the service performance of the dies.

Apart from conventional hardfacing alloys, weld material could come from one of

the many advanced materials, like ceramics and intermetallics. Ceramics are chemically

inert compounds that can retain its properties at high temperatures. These may be

potentially good candidates for applications where toughness is not an issues but hot

strength and resistance to oxidation is very critical. Hot pressed Silicon Nitride is a

ceramic that has extremely high hardness, high toughness and wear resistance. Currently it

is used in applications like nozzle rings, bearings, rotors and cam followers in internal

combustion engines. It also has good thermal shock resistance and good hot hardness and

maintains its temperature and oxidation resistance at 1200° C. Applications of Silicon

Nitride as a coating, is however, limited because of its poor adhesion with the substrate.

Silicon Aluminum Oxynitride (Sialon) is a new group of solid solution

compositions that also possesses excellent thermal shock resistance. Sialons have similar

properties as Silicon Nitride but Sialons have a superior resistance to oxidation at high

temperatures. Silicon Carbide with extremely high hardness is normally used in grinding

wheels as well as various internal combustion engine parts like valve seats and flame cans.

Several ceramics and carbides exist that have application in precision forging either as

58
inserts or as coatings. It should be realized that these compounds typically lack tensile

strength and needs to be constrained in a shrink ring. Also, they may not be applicable as

coatings because of their lack of adhesion to the substrate and their dissimilar thermal

expansion coefficients that may lead to cracking.

Nickel aluminides are relatively new intermetallic die materials that exhibit better

high temperature properties compared to conventional hot work steels and nickel based

superalloys like IN 718. Although this is a relatively new compound, it is the same

compound (γ’ compound) that gives superalloys like 718, its strength on aging.

The application of Nickel Aluminide in hot and warm forging is very new.

Although soft at room temperatures, nickel aluminides retain their yield strengths at higher

temperatures. Figure 2.18 shows typical yield strengths and tensile strengths of Nickel

Aluminides compared to hastealloy and stainless steel. . The yield strength of these

intermetallic compounds increases with temperature up to 500 °C beyond which the yield

strength drops. The relatively high hot hardness gives these intermetallics, very high wear

coefficients. Tests that have been done industries show up to 10X life increase for

preforming dies. The high yield strength also gives nickel aluminide relatively late crack

initiation as indicated.

59
Figure 2.18. Comparison of yield strength of IC-15 to those of other high temperature

alloys.

2.4 Conclusions

The yield strength of the die material at the surface, exposed to the high contact

temperatures, is one fundamental property influencing die failure. It affects the occurrence

of the following failures:

• Plastic deformation

• Thermal fatigue (crack initiation)

• Mechanical Fatigue (crack initiation)

60
• Wear

As the hot and warm forging dies reach high temperature during the working

cycles, it is necessary that the hot yield strength stays stable during the hot work. The yield

strength is the property that directly resists the working pressure, and keeps the dies

working in the elastic field macroscopically. This working condition will provide forged

pieces inside the geometrical tolerance range. However, critical regions of dies can be

subjected to high stresses that can lead to plastic deformation. The thermal stress-strain, or

the mechanical stress-strain state can cause thermal fatigue or mechanical fatigue,

respectively. As the amount of plastic strain is the driven cause for fatigue crack initiation,

high yield strength will reduce the amount of plastic strain, retarding or avoiding the crack

initiation. The wear resistance is direct proportional to the yield strength, represented in

the models by the hardness.

The ductility or plasticity of the hot work tool steel is other important property.

Although the dies should work in the elastic regime, localized plastic deformations can

occur. The plastic deformations from thermal origin are difficult to avoid, especially in

regions of high thermal load. In this case, the number of cycle to crack initiation will be

direct proportional to the ductility limit (area reduction in tensile test). Also, crack

propagation is believed to be controlled by plastic deformation in the low cycle fatigue

regimes typical in warm and hot forging.

The toughness or fracture toughness is the materials ability to resist crack growth.

This property will allow the dies to work at a higher stress-crack size without reach the

61
condition for fragile fracture that will leads to the die catastrophic failure. Low toughness

also increases the crack growth rate.

From the fundamentals of wear, we can safely say that apart from hardness and

subsequent softening of die materials, pressures and the amount of sliding also affect the

failure rates. As discussed in previous chapters, pressure is primarily dictated by forging

material, forging temperature, lubricant used and the geometry of the dies. It also depends

on the die closure, flash and the type of forging equipment used. The preform shape,

lubricant and flash, control sliding distances dies experience during forging. Surface

hardness depends on, apart from the alloy composition and microstructure, coating or

surface treatments used, lubricant, thermal cycling and to an extent, preheating.

Unfortunately, the interaction of the forging parameters and the wear and failure

rates is too complex to draw any direct correlation. For instance, forging temperature

reduces the wear resistance of the surface. Also, higher temperatures typically produce

thicker, but not necessarily more adherent, scale. Scales, if adherent, increases wear.

However, the loads felt by the die are lower because of the lower flow stress of the

material at high temperatures. Wear rates, here, would be controlled the relative

magnitudes of these effects and can be predicted only by analyzing all these factors

together.

It is also important to note that dwell times, heating times and cycle times also

have contrasting effects on wear of dies. Increased heating time would increase scale

62
formation. In a 2 or 3-step operation involving descaling, this would not be a big factor.

But process designs that employ single blows should pay special attention to heating times

and heating atmosphere. On a side note, thicker scales act as insulators and keep the billet

hotter. Sometimes, scales also act as lubricant, reducing loads. By increasing dwell time,

when the die is in contact with the billet, the lower dies experience substantial softening.

But this also cools the billet in contact with the lower die, reducing sliding. Preheating,

though effective in reducing the chance of catastrophic failure and thermal fatigue,

increases wear by reducing the hardness and wear resistance.

As we see, there are many controlling factors that affect wear and die failure. It is

important to evaluate wear as a cumulative result of all these process variables. There are

several models proposed in the literature that try to capture some of these relationships.

There are several interrelated parameters that affect the performance of forging

dies. In working to improve hot and warm forging dies performance the fundamental step

is to identify what is the dominant failure mechanism. Only with this information, it is

possible to improve the correct properties, and optimize the correct process parameters

that will result in better die performance. During this process, it should be noted that

solutions to reduce wear are different from those that reduce thermal fatigue and

mechanical fatigue. Use of computer simulations [Painter, et al., 1996, Tulsyan, et al.,

1993] could be a necessary first step to evaluate the conditions at the die-billet interface

before a good solution can be obtained.

63
Although there are several modes of die and tooling failure, because of safety

reasons, the main concern of a tool designer is catastrophic failure of dies. In very few

cases, designers employ predictive model to design tooling to avoid catastrophic failure.

The tendency, in forging industry, is to use material with low hardness and high toughness.

However, beyond a point, toughness does not bring any benefit to the die life. Because of

lock of good understanding of fatigue failure, the design and material choice is done very

conservatively.

Low hardness, because of lower alloying content, reduces the wear resistance. The

wear resistance is function of the tool steel hot hardness, and the carbides in the matrix

(amount, size, and hot hardness). However, generally these carbides reduce the ductility

and toughness. Carbides, necessary to resist wear, can be detrimental to resist die fracture

(toughness) and fatigue (ductility). The wear resistance needs the material to possess hot

yield strength. Thermal fatigue resistance is improved by both critical ductility is at room

temperature and hot hardness. The alloying contents command both hot hardness and

ductility. By carefully tailoring the microstructure and alloy content to the application, it is

possible to balance the different failure mechanisms such that the tool life is highest. The

alloying content in the matrix can be modified by the heat treatment that controls the

dissolved alloys in carbide form. However, higher the undissolved carbides, higher is the

wear resistance. But the thermal fatigue resistance reduces with higher amounts of

undissolved carbide. As can be seen, there is a complex inter-relation between failures

64
mechanisms and properties that need to be understood and applied correctly to improve

the life of the dies for hot and warm forging.

We understand that the wear needs a more detailed evaluation under the conditions

usually found in hot and warm forging, and the same is valid for the interactions between

wear and the thermal fatigue. Based in these needs we developed a new test for

applications to die forging at high temperature that can evaluate simultaneously the wear-

thermal fatigue failure mechanisms.

It is essential to understand the mechanisms of die failure completely before we can

attempt to increase die lives. As mentioned before, die wear is the major mechanism of die

failure in high temperature forging, followed by mechanical fatigue. We emphasize that the

wear failure initiation can caused or increased by thermal fatigue, as indicated by: micro

observation of the die cavity, and the higher wear rate in dies with more severe

temperature cycle. However, in most cases, several modes of failure act in conjunction. In

this section, we summarize the effect of various criterions on die failure.

65
CHAPTER 3

MODELING OF TEMPERING AND WEAR

3.1 Overview of Past Studies in Modeling of Tempering, Wear and Die Life

It is clear from the previous section that wear is the primary mode of failure in hot

forging. This section will go into the physical and metallurgical phenomena occurring at

the surface and list a few process issues that could positively or negatively impact wear.

Figure 3.1 shows various factors that control extent of abrasive wear in hot forging.

Abrasive wear in hot forging occurs due to the following series of events.

• The die surface is heated to extremely high temperatures either because of high

contact times or due to poor or no lubrication.

• Prolonged and repeated exposure to high temperatures results in oxidation of surface

and gradual loss of hardness at the die surface

At convex areas of the dies, metal slides past radii at pressures ranging from as low as the

flow stress of the metal (near the flash region) or as high as 5-8 times the flow stress of the

66
material near the centers. Hard particles in the forging material, scales and other hard

contaminants erode the soft shell of the dies gradually degrading the die surface.

Figure 3.1. Various process and material

Figure 3.2 clearly shows how the temperatures of dies vary during a forging

process for a hot forging performed in a hydraulic press and a hammer. In a hydraulic

press, the dies are exposed to hot billets for a longer period of time compared to hammer

67
forging. Subsequently, dies could get substantially hotter compared to hammer forging

dies. Also, it should be noted that because the lower dies experience higher temperatures

compared to the top die because the billets are in contact with the bottom die for a longer

period of time.

Figure 3.2. Variation of die temperatures during a hot forging cycle

68
Also, from the event progression, the following can be deduced

1. Higher preheat temperatures will result in higher overall die temperatures and

hence a higher hardness loss at the surface. For most tool steels, this translates into

a higher rate of wear (Figure 3.3). Die preheating also affects phase transformation

at die surface. In some cases, the phase at the surface transforms from martensite

to austenite during the heating followed by the transformation back to untempered

martensite [Okell and Wolstencroft, 1968]

2. Higher forging temperature typically increase the die surface temperatures and

possible oxidation of the die surface

3. Better lubrication reduces heat transfer into the dies and will reduce extent of

tempering that occurs at the surface. Also, typically, better lubrication also lowers

the local pressures on the tool surface that causes wear rates to increase (Figure

3.4)

69
Figure 3.3. Effect of forging temperature on the wear depth after forging 4000 pieces

[Netthofel, 1965]

70
Figure 3.4. Variation of wear volume with die bulk temperature for lubricated and dry

forging [Singh, et al., 1973]

Metallurgical changes on the die surface

While the bulk die temperature in hot forging is usually around 350°F, the dies

sub-surface temperature usually reaches 1100°F, although the peak temperature at surface

can reach as high as 1650°F in certain applications.

To understand the effect of tempering temperature and time on hardness, one

needs to understand the transformed microstructure of H-13 or other tool steels. Wear

resistance and resistance to plastic deformation and thermal fatigue depends directly on

the structure. The following section describes the changes that occur in tool steels during

71
the heat treatment process. It is necessary to understand the heat treatment process before

one understand what happens to tools exposure to prolonged high temperature during

service.

Quenched structure of any tool steel usually is a combination of untempered

martensite, retained austenite and carbides retained from austenitization. Typical hot

forging tool steels like H13 transforms into martensite completely if it is quenched to

300°C in 1000 seconds. Martensitic state is formed by sudden shear of parent austenite

because the rapid cooling that results in martensite does not provide any time for diffusion

of carbon atom from the austenitic lattice. Given the opportunity, carbon and other

elements will diffuse towards forming spherodized structures that represent lower energy

states.

Low alloy or plain carbon steels rapidly loose their hardness when tempered. This

is primarily due to rapid coarsening of cementite. For carbon steels, tempering occurs in

well documented 3 stages. These stages are

• Between 200-400 °F transition carbides are formed and the carbon content of

martensite is lowered to about .25%

• Between 400 and 575 °F retained austenite is transformed to more stable bainite,

ferrite and cementite. This usually results in substantial softening of the structure.

Above 500°F, the matrix slowly transforms to ferrite and cementite, which grows in size.

In tool steels, these mechanisms are slowed down due to the presence of alloy elements.

72
Here, the first stage of softening (2-6 Rockwell points) occurs from room

temperature to about 520 °F. In this stage, the martensite decomposes to tempered cubic

martensite accompanied by the rejection of carbon from the matrix. This forms finely

divided epsilon carbides, a transition phase that disappears as cementite forms between

570 °F to 750 °F. With time, cementite formed also starts growing in size. This growth is

however, much slower than plain carbon steel because of the slower rates of diffusion.

During the second stage of tempering, occurring between 750 °F and 1050 °F,

cementite partly goes into solution. This stage is accompanied by precipitation of

secondary alloy carbides. In H-13 and other medium to high alloy tool steels, diffusion of

carbon is substantially slowed by the presence of alloy elements. Alloy elements diffuse

substitutionally unlike carbon, which diffuse interstitially. As a result, these reactions

require substantially longer exposure to high temperatures. Because of precipitation of

metallic carbides, there is an accentuated reduction in toughness in the region of secondary

hardness peak. This is illustrated in Figure 3.5. For hot forging dies, it is important to

temper at a temperature that exceeds this critical point. Tempering at temperatures just

above the secondary peak gives the maximum hardness, desirable for good performance in

thermal fatigue and wear. However, if the dies experience extreme mechanical loads, an

increase in tempering temperature increases the toughness.

73
(a) (b)

Figure 3.5. Drop in toughness of H-13 with secondary hardening. a) represents toughness

in Charpy V-notch energy, b) represents toughness in Kic [Pickering, 1987]

The third stage of tempering occurs when retained austenite transforms to

secondary bainitic and martensitic structures resulting in marked increase in hardness. H-

13 is typically double tempered and higher alloy tool steels triple tempered to ensure

retained austenite content is reduced. When tools are put into service they have are usually

heat treated to a point beyond the secondary harness peak. This also ensures that the

toughness never gets any worse during service.

During service, subsequent exposure to elevated temperatures results in coarsening

of both secondary alloy carbides and cementite and the gradual depletion of carbon from

the martensitic matrix. This gradually moves the structure at the surface down the

hardness scale. However, the core properties do not deviate substantially from the initial

set of properties.

74
Alloying elements have a strong effect on the die’s resistance to thermal softening.

Higher carbon results in higher hardness achievable. However, hardenability is unaffected

by carbon content. As a result while the surface hardens due to quenching and tempering,

the core remains relatively soft. Also, more of the surface austenite transforms to

martensite, which is less dense than ferrite-pearlite mix. The phase differential causes

quench cracking. Also, carbon dominated hardness quickly tempers on exposure to high

temperatures as the diffusivity of carbon is very high.

Tool steels in hot work applications rely on strong carbide formers, like V, Mo, T,

Cr, etc. to provide hardenability and resistance to thermal softening. Not only do the alloy

carbides provide hard particles to resist wear, but the presence of alloy atoms in the crystal

structure slows down diffusion of carbon.

For example, .13% Vanadium has the ability to flatten the slope of a tempering

curve by 50%. This means that after 2 hours of tempering (after quenching or during hot

working), tool steel with .13% V will result in a drop of hardness of half the value

compared to a pure iron-carbon structure. However, in order for this to happen, these

elements have to be dissolved in the austenite during austenitization.

75
3.2. Characterization of Thermal Softening

Hardness of as-quenched martensite is a function of the carbon content. Tempering

of this structure improves the strength and toughness of the quenched die steel by stress

relieving and atomic rearrangement. Due to reasons discussed before, the tempering time

and temperatures have similar effects on the microstructure, and is heavily influenced by

the material structure. Modeling this phenomena through a phenomenological approach, is

hence extremely difficult and impractical. Hence manufacturers resort to empirical curves

to indicate the tempering characteristics of materials.

Usually, the tempering curves from steel producers provide variation of hardness

with tempering temperatures for a fixed time of one or two hours. Heat treaters use this

information to come up with charts and thumb-rules that relate the tempering times to the

bar or tool thickness and the hardness required. No fundamental models have been

reported that relates hardness measured to softening mechanisms that include time and

temperature.

To reveal the tempering effect, Shi [Shi and Liu, 2004] performed tensile tests at

elevated temperatures with various soaking times and heat treatment. He performed these

tests on hardened AISI 52100 steel. Shi found that at the same temperature level the flow

stress is smaller for the specimens that received stronger tempering, and the decrease of

strength due to tempering effect becomes more pronounced at high temperatures. He

76
modeled tempering effect and expressed it as a function of thermal history and material

hardness.

A better approach to specifying effect of tempering is using charts and functions

that track hardness with a combination of time and temperature. The first relationships

that relate time and temperature to the resulting hardness were proposed by Hollomon and

Jaffe. Several work followed this initial effort. Figure 3.6 shows the master tempering

curves for H-13 using parameters by Larson and Miller and Holloman and Jaffe.

These curves provide us a means for calculating the effects of short spurts of high

temperatures on the hardness, by giving us a method to compare and accumulate relative

effectiveness of various levels of temperatures and times. The Hollomon Jaffe parameter

(or an equivalent Larson Miller parameter) captures the dominant influence of temperature

in relation to time, using a model of the form given in parameter equation 3.1.

Master tempering parameter = T . (C + LOG (t)) (3.1)

where C is a material constant, T the temperature and t is the isothermal time at

temperature T. Plots of the master tempering parameter with the hardness serve as a

means Using these curves, we can predict what the resulting hardness will be if we temper

a steel block for a specific duration at a specified temperature. For steels, C is generally

between 10 and 15. For H13 tool steel, the tempering parameter is about 16.5 (and 20 if

Larson Miller relationship is used).

77
According to Oddy, the degree of tempering can be defined by a dimensionless

ratio in terms of material hardness change, as

H0 − H
φ= (3.2)
H0 − H∞

where, H(0) is the initial, as-quenched hardness of the work material, H(∞) is the fully

softened hardness value, H is the tempered hardness at moment t. While a simplified

linear relationship between the tempering parameter (Holloman Jaffe) and the hardness

works under most cases, Oddy [Oddy, 1996] proposed a exponential relationship

(Equation 3.3)

 Tempering.Parameter 
φ = A. exp  (3.3)
 B 
where A and B are material constants.

3.3. Prior work in Wear Characterization

There are several trends and thumb-rules that exist that describe the effect of

various process parameters like lubricant application and forging temperature on wear.

However, any trend found in the literature, derived from actual production setting, has to

be treated with caution. Information obtained from production source is “noisy” because

there is little control over forging process variables like temperature, forging cycle times,

lubricant spray duration and quantity. In addition, data and trend applicable to one part

78
family may not be applicable to a different part family. Differences in process conditions,

die and forging material may result in a completely different wear results.

Figure 3.6. Master tempering curve for premium H13 were: P= Larson-Miller parameter,

T is temperature (°F), t is time in seconds [Carpenter, 1999]

As discussed earlier, wear in hot and warm forging is directly proportional to the

interface pressure, sliding distance and inversely proportional to the die’s surface hardness.

If geometry of the die is complex, modeling and predicting the sliding distances and

pressures is difficult. Quantifying wear in such cases can also be tedious. Hence, several

researchers have resorted to using simple geometries for testing.

Controlled experiments to overcome the deficiencies of modeling with production

data have been performed with simple cylindrical work pieces [Aston and Barry, 1972,

79
Aston, et al., 1972, Batit, et al., 1983, Dennis and Mahmoud, 1987, Felder and Montagut,

1980, Felder and Montagut, 1980, Sharma and Arrowsmith, 1981, Thomas, 1970,

Tittagala, et al., 1982], extrusion type pieces [Doege, et al., 1996] and more complex

geometries[Dean, 1974, Dean, 1966, Silva and Dean, 1971]. A few of these pertaining to

precision forging are reviewed in this paper.

Wear evaluation performed by forging cylindrical test billets between flat dies

relies on measuring wear profiles using surface finish measuring machines. The test setup

used and typical profile obtained after substantial wear of the flat dies is shown in Figure

3.7.

Figure 3.7 Typical cylinder tests used to characterize wear and resulting measurable wear

pattern [Singh, et al., 1973]

80
Tittagala and Bramley [Tittagala, et al., 1982] recommended and performed pin-

on-disc type experiments to study wear of different die materials. The schematics of their

tests are shown in Figure 3.8. In this setup, they mount the die in a lathe spindle and the

heated workpiece in an eccentrically positioned tailstock. The setup has provisions to

exert an axial pressure on the die and intermittently heat the dies and the workpiece to

predetermined set points. However, Bramley and others conduct the test at an axial

pressure lower than the yield strength of the billet material. After the test setup was

validated, Bramley [Bramley, et al., 1989] conducted tests to study the effect of

workpiece temperature, die preheat temperature, load, contact duration, frequency of

lubrication and lubricant concentration on the wear.

Figure 3.8. Experimental setup used by Tittagala and others [Tittagala, et al., 1982]

81
More recently, researchers around the world are trying to combine FEM

techniques to design more complicated tooling for wear testing. Bariani and other

[Bariani, et al., 1996] have used a Gleeble 2000 testing machine to upset cylindrical steel

(DIN Ck40) billets between double acting punches made of X38CrMoV51 heat treated to

54 Rc. The punches had more complicated geometries (Figure 3.9). Using this setup, they

were able to observe and measure wear using a coordinate measuring machine after only

forging 50 pieces. Based on wear measurements, tool hardness measurements and

pressure and sliding velocities obtained from layered wax models and finite element

simulations, they were able to arrive at wear constants for different regions of the punch.

Silva and Dean [Silva and Dean, 1971] also conducted trials using dies that

resembled real production dies (Figure 3.10). They used cold rolled EN-8 billets induction

heated to 1040° C and forged on a Mk IID Petro-Forge hammer. The billets were not

descaled before forging. The dies were made of Uddeholm’s 5% Cr die steel, similar in

composition to H-13, heat treated to 46-48 Rc. Doege, Melching and Kowallick [Doege,

1978] used automated heading machine to forge 1045 steel disks into cups using an

extrusion type process (Figure 3.11). Die heaters were used to keep the temperature

within preset range. After every 300 pieces, Doege and coworkers used a surface-

measuring machine to measure the average distance between the worn and unworn profile.

They used this as a measure of wear.

82
Figure 3.9. Punch designs used by Bariani [Bariani, et al., 1996]

Figure 3.10. Dies used by Silva and Dean for wear characterization [Silva and Dean,

1971]

83
`

Figure 3.11 Extrusion – type tests conducted by Doege, Melching and Kowallick [Doege,

1978].

Doege, [Doege, 1994] also performed tests with forging dies that resembled dies

used in actual production. The press used in his tests was a Eumuco eccentric press. The

tests utilized dies made of H10, H12 and L6. The dies were preheated to 220° C and the

billets were inductively heated to 1120 °C. Lubricant used was a water based graphite

suspension (Delta 31). The objectives of his tests were to see the effect of temperature

84
and water cooling of dies on the wear. He measured wear as a geometric deviation at the

punch cone radius indicated in Figure 3.12.

Netthofel [Netthofel, 1965]used a friction screw press in his studies to evaluate

wear. His forging speeds were .5 m/sec. The forging stock he used was .53 percent carbon

steel (.53% C, .26% Si, .7% Mn, .02% P, .02& S). A gas flame heater controlled the die

temperatures to a set point. The forging cycle time was kept at 13 seconds. Round corner

square billets were heated to 1200° and de-scaled in a mechanical press before they are

forged in the screw press. The setup Netthofel used is shown in Figure 3.13. Netthofel

used sawdust as a lubricant that he applied before every blow. Die wear was measured in

the locations indicated in Figure 3.13. Netthofel employed pin shaped inserts that were 16

mm in diameter and made from different die materials. By drilling holes along the

circumference of the die, he was able to subject the pins to substantial sliding and

pressure.

Using these tests, researchers have been able to compare the wear resistance of

various die materials, surface treatments and coatings. However, extending these to

applications outside the domains the tests were done, does not give reliable results.

85
Figure 3.12. Doege’s experimental setup [Doege, 1994]

Figure 3.13. Test setup used by Netthofel [Netthofel, 1965]

86
CHAPTER 4

FORMULATION FOR PREDICTING WEAR AND THERMAL SOFTENING

4.1. Proposed Formulation for Thermal Softening and Wear of Hot Forging Tools

To overcome shortcomings of the existing wear and softening model and to

correlate the failure characteristics of tools to the process conditions, we need

fundamental / phenomenological models that captures

• Die softening or over tempering

• Abrasive and Oxidation wear

Thermal fatigue

These three mechanisms play crucial roles in failure of hot and warm forging

applications. This chapter presents the tempering model and wears model developed along

with methodology to integrate these into FEM.

87
4.1.1 Intermittent Tempering Model Formulation

As evident, the alloying elements have a strong effect on the dies resistance to

thermal softening. It is also obvious from the prior section that tempering of tool steels is a

complex diffusion controlled phenomena that is both time and temperature dependent.

A subroutine that calculates effective tempering that occurs during a forging cycle

has been designed and implemented. The subroutine works with DEFORM-2D, a explicit

special purpose finite element for modeling bulk deformation. DEFORM is widely used in

forging industry in US and is a suitable platform for application being tried out. A

schematic representation of the algorithm used in designing the subroutine is shown in

Figure 3.7. Calculations to be performed at the end of the computation step (after the die

temperatures have been calculated). All tempering temperatures have been normalized to

600 °C or 873 K.

88
nth time step

Initialize variables

mth node
Read in die
material

Tempering
Acquire Holloman
Material
Jaffe constants
database
from database
∆t(n),
Calculate effective T(m-1),
tempering in δt T(m),
T(ref)

Accumulate
tempering time

Calculate new HJ
parameter

Calculate hardness

Advance to next Advance to next


node node

Figure 4.1. Schematics of modeling in-process hardness loss of forging tools

89
4.1.2 Architecture and Model Development

The following model is used in calculating loss of hardness of dies during thermo-

mechanical processing of materials. The equations used are

  Ti (t ) + T f (t )  
T =    + 273
 (4.2)
 2  

T ( C + log δti )
( −C )
δt ET = 10 T ( ref )
(4.3)

TET = TET + δt ET (4.4)

H = 1273× (C + log(TET )) (4.5)

where δti is the incremental time step in seconds, Ti(t) is the temperature at the start of the

computational step in C, Tf(t) is the temperature at the end of the computation step in C,

T(ref) is the reference temperature in Kelvin, T is the average temperature (in Kelvin)

during the time step, C is a material constant (user input, material dependent) and H is the

Hollomon-Jaffe parameter. A look-up table is needed to translate this into hardness loss.

4.1.3. Formulation of Wear Model

As outlined before, several works exists in the literature that tries to characterize

and model wear in hot forging. Some of the earlier work focused on characterizing wear

as a function of process variables like forging area, weights and energy while some have

taken a more fundamental approach to modeling. These models are provided in Table 4.1.

With advances in finite element models and computing, it is possible to use fundamental

90
material properties and process variables derived from FEM software to model wear more

universally. With the technological capabilities in mind, and with available data, it is

possible to use Archard’s model provided in equation 4.3 to model wear as a function of

thermo-mechanical history of dies during a forging process and the working hardness of

the die material.

p i × Vi
wear = k ∫ dt
Hi (4.3)

where p is normal pressure at a die location, V is the sliding velocity at any time, H is

the hardness of the die location and k is a constant dependent on several factors like billet

material and scale formation.

Earlier models that relied on tests to characterize wear coefficients are not

fundamentally sound. Although these tests give a rough measure of wear resistance for the

specific class of forgings, these cannot be extended to forgings outside the class.

Phenomenological models like Archard’s model are far more extendable to other

applications, provided the mechanism of wear is similar. For instance, models developed

or calibrated for abrasive wear can be used for other abrasive wear applications. However,

these cannot be used where adhesive and oxidation wear dominates abrasive wear. Also,

Archard’s model does not take into account the varying softening behavior of different

materials, with exposure to high temperature. Low alloy/low carbon steels have a steep

drop-off in its strength/hardness with temperature compared to high alloy tool steels. In

fact, some materials (like intermetallics) get stronger at higher temperatures.

91
Aston and Hammer
Barry Mean damage (x10-3) = 0.00686 x forging area + 0.0272 hammer energy -
{Aston, 1972 0.1855 x forging wt1/3 + 0.335 x spread - 0.011 x flash land area +
#247} 0.129 x flash metal escape – 0.557

Aston and Hammer


Barry Mean damage (x10-3) = 0.000261 x forging area + 0.000763 hammer
{Aston, 1972 energy - 0.00265 x forging wt1/3 + 0.012 x spread ration - 0.000694 x
#247} flash land area – 0.00266

Aston and Press


Barry Mean damage (x10-3) = 0.0284 x forging weight – 0.062 x die material - +
{Aston, 1972 0.141 (83%)
#247}
Aston and Mean damage (x10-3) = 0.000164 forging area + 0.000712 x flash
Barry land/gap - 0.00431
{Aston, 1972
#247}
Archard’s pi × Vi
model wear = k ∫ dt for volume
Hi
k = constant,
p = normal pressure
V = velocity
H = hardness

−4
Budinski w = 0.23 × 10−2 e − 0.21 x10 A
{Budinski,
1980 #316} w = abrasion rate cm3/min
A = Structure parameter for a given tool steel (carbide size (µm) x volume
fraction x carbide hardness (kg/mm2))

Thomas R = 204 - (70 (%C) - 4 ( %Si) - 15 (%Cr)1/2 - 80 (Mo*)1/3


{Thomas, Mo* = %Mo + 0.5%V + 2%V + %Nb
1970 #321} R is the wear rate relative to H13 steel

(*) Model used by Bariani {Bariani, 1996 #342}, Batit {Batit, 1983 #314} includes an exponent to the
hardness Hm, Painter {Painter, 1996 #285} uses Archard model with computer simulation.

Table 4.1. Various wear models reported in the literature

92
Also, Archard’s model does not take into effect the in-service tempering behavior

of tool material. The surface hardness of dies exposed to prolonged high temperatures is

substantially lower that the core hardness. Also, surface also experiences decarburization.

These phenomena make the surface more vulnerable to wear – a mechanism not captured

by the simple Archard’s relationship suggested earlier.

With advent of robust FEM engines, one can now embed these mechanisms into

the simplified Archard’s model, to extend the model across various processes. Similar

effort is being carried out by Kang [Kang, et al., 1999, Kang, et al., 1999], Jeong[Jeong,

et al., 2001]and Saiki [Saiki, et al., 2001].

4.2. Demonstration of Thermal Softening Model

Tempering of the surface layers of the forging tools greatly affects the wear

resistance of the tool. Process parameters like the equipment choice, forging temperature,

lubricant quantity and grade and cycle times greatly affect the thermal history the dies are

subjected to. These are well known among the research community, and to some extent

understood by the forgers. However, the relative effects of these on the magnitude of

tempering and hence wear, have not been looked at, in its complexity. The following

section intends to demonstrate, through the use of the developed models, these effects, on

the dies used for hot forging a gear blank. Design of tools and process details have been

obtained from Sypris Technologies, Forge Division in Marion, OH.

93
The modeled process forges a gear blank from AISI 8620, 2” diameter and 4”

long. Parts are being forged on a 1600 ton mechanical press with a total stoke of 14” and

a speed of 60 spm. Forging temperature was 1100 °C and the forging cycle time was 15

seconds. Between successive parts, the dies are lubricated manually with Acheson’s

Deltaforge F-31, a fine water based colloidal graphite. The forging is being done in 2

passes: the first pass pancakes the billet to remove scale and oxide layer that is forms

during heating, the second pass forges the pancake to the finish size. Dies used were made

of H-13, quenched and double tempered to a hardness of 48 to 50 HRC (Rockwell C

scale). Before operation, these dies are typically preheated to 200 °C for better fracture

toughness. A schematic representation of the 2 pass forging is shown in Figure 4.2.

These conditions have been used to model the process to establish a benchmark for

prediction of material over-tempering and wear.

4.2.1. Effect of lubricant heat

To evaluate effectiveness of forging lubricants on tempering, wear and tool

degradation, several simulations were performed with varying heat transfer coefficients.

The values of heat transfer coefficients used in the computer simulation corresponds to the

values obtained and in some cases, extrapolated from the tests performed at the Ohio

State University by Sridhar and others. Values of the interface heat transfer coefficients

used were 12 KW/m2°C, 24 KW/m2°C and 33 KW/m2°C, corresponding to a water based

graphite lubricant diluted with 20 parts, 30 parts and 100 parts water (dilution ratio of

94
1:20, 1:30 and 1:100). The hardness distribution of the top die after 3000 and 5000 shots

for each of these dilution ratios, is given in Figures 4.3, 4.4 and 4.5. The hardness

distribution along the surface for a dilution ratio of 1:20, is shown in Figure 4.6.

Figure 4.2. 2 stage gear blank forging sequence (Courtesy: Sypris Technologies)

95
Figure 4.3. Hardness distribution after 3000 and 5000 shots, heat transfer coefficient used

= 12 KW/m2°C, press type: mechanical press

Figure 4.4. Hardness distribution after 3000 and 5000 shots, heat transfer coefficient used

= 24 KW/m2°C, press type: mechanical press

96
Figure 4.5. Hardness loss of the top die nose and flash areas after 3000 and 5000 strokes,

assuming a heat transfer coefficient of 33 KW/m2°C

97
Figure 4.6. Hardness distribution at the surface, after 3000 and 5000 shots, heat transfer

coefficient used = 12 KW/m2°C, press type: mechanical press

From the above plots, the following conclusions can be made.

• Regions in the center are in general, exhibit a more prominent tempering and hardness

loss because these regions come in contact with the billet earliest

• As Figures 4.3 and 4.6 indicate, corners have a more pronounced hardness loss. This

may be due to the lower mass content in corners (higher surface to volume ratio) that

disproportionately increases the area through which heat is conducted in. Typically the

corners are the regions with the highest tangential or sliding velocities. Coupled with

the large hardness loss, corners towards the center of the part, could possibly wear the

fastest.

98
The fillets exhibit the least hardness drop because these are the last areas to fill and come

in touch with the hot work piece material

In these computer simulations, heat transfer coefficients are assumed to be

independent of pressure. However, some research exists, that show that the lubricant film

quality changes substantially with the local pressure and temperature. This in turn will

affect the local friction and heat transfer coefficients, which in turn will have a dramatic

effect on the local over tempering of the tool surfaces.

4.2.2. Effect of Equipment

To demonstrate the effect of equipment selected and contact time, several

simulations have been performed with mechanical press and hydraulic presses with various

ram speeds. By changing the press characteristics, we are able to change the duration in

which the tools are in contact with the workpiece. In these simulations, the heat transfer

coefficient has been kept constant in order to isolate the effect of equipment, on the

hardness drop.

Figure 4.7 and 4.8 show hardness distribution in the top punch, if these forgings

were made on hydraulic press at a forging speed of 10 mm/sec and 20 mm/sec. The

friction factor used in these simulations was 0.3 and the heat transfer coefficient used was

6 KW/m2°C.

99
Figure 4.7. Hardness distribution after 3000 and 5000 blows, with a slow hydraulic press

(p2) v=10 mm/sec, h = 6 KW/m2 °C, m=0.3

It was noted that the softened surface effect was much more predominant in a

hydraulic press compared to the mechanical press due to the longer contact times typically

found in hydraulic press forgings.

100
(a) (b)

Figure 4.8. Hardness loss after (a) 3000 and (b) 5000 blows, with forging on a hydraulic

press with ram speed v=20 mm/sec, h = 6 KW/m2 °C, m = 0.3

4.3 Summary

This chapter presented some background on tempering phenomena in dies used in

elevated temperature forging. It also presented the formulation of the finite element model

for non-isothermal simulation of die tempering processes. Empirical relationships used and

how these were integrated into the finite element code and its use demonstrated using a

industrial case study involving forging of a idler gear used in truck transmission.

101
CHAPTER 5

VALIDATION OF WEAR AND OVER-TEMPERING

5.1 Introduction

Chapter 4 outlined the algorithm developed to predict the effects of the thermal

cycling, sliding and pressures on the thermal softening of the die subsurface and wear. This

algorithm outlines the various inputs to the model and the model outputs. Inputs to the

thermal softening model are the tempering parameter of the tool material, the initial

hardness and the process information. Inputs to the wear model are the material hardness

(at room temperature and elevated temperatures) and the wear coefficient. Another

product based input is the tolerances acceptable. The outputs are the tempered layer

hardness profile, wear and tool life. The following section describes test setups that are

used to validate the thermal softening and wear models.

5.2 Thermal Softening Setup

The thermal softening algorithm described in the previous chapter employs finite

element approach to calculate effective tempering at various die locations based on

tempering parameters similar to ones proposed by Hollomon and Jaffe and modified by

102
Grange and Baughman. After the thermal profile is computed by the FEM engine, the

developed module uses the time step information to calculate the net tempering of the

various die location in the time step, using the relationship 4.3. During the simulation,

these incremental “tempering times” are added up to calculate the net tempering time the

different die locations are tempered for at a reference temperature.

Based on the characteristics of the die material, these effective tempering times are

translated to hardness loss. For most die materials, the tempering parameters outlined are

published. For instance, for H13, the Grange Baughman tempering parameter is 16.44. If

the tempering parameters are not easily available, one may be able to calculate these based

on isothermal tempering data available as part of the material data sheets.

The thermal tempering model was validated using a test setup that recreates a

intermittent thermal cycle seen in forging. During a hot forging cycle, the billet sits on the

lower die for duration of 2-3 seconds. This is followed by the forging blow. During this

period, the heat transfer into the dies goes up, because the tools come in intimate contact

with the hot billet. During this phase the subsurface temperatures could reach in excess of

500 °C, while the bulk of the die remains at the steady state temperature of 200-300 °C.

The surface and subsurface temperatures vary depending on the process, the type of press

used, the billet temperature, the die bulk temperature and other factors like the presence of

scale. Also, features in the die like sharp corners, gear and spline teeth and notches tend to

over heat resulting in hot spots. If not careful, die surface temperatures may even exceed

the steel transformation temperature. After forging, the part is ejected and removed and

103
lubricant sprayed. Surface cools down sharply, while the substrate remains at a relatively

high temperature (close to the steady state temperatures). The next billet is then placed on

the bottom die and the cycle continues. Typical thermal cycle found in a hydraulic press

and a hammer forging die is shown in Figure 5.1. The thermal spike typically seen at the

surface of these dies is illustrated in Figure 5.2.

Figure 5.1 Typical thermal cycle in a hydraulic and hammer forging

104
M ax T = 1000 - 1200 F

Distance from surface

Figure 5.2. Temperatures spike seen at the die surface, in a typical hot forging die.

Simulation was performed in DEFORM2D.

To study the effect of this thermal cycle on the subsurface hardness, a simple test

setup was used (schematic shown in Figure 5.3). The test relied on cyclically contacting

and retracting a die material test specimen onto and from a hot billet. By controlling the

contact time, the pressures at the interface and the time the sample cools, thermal cycles

very close to a hydraulic press forging was achieved.

For this test setup, the billet material was made of Waspaloy. This choice was

made so that the billet material did not deform during the forging cycle. Also, this choice

ensured that the billet does not oxidize over the duration of the test. These steps were

intended to reduce the variability between stroke to stroke, to the maximum extent

105
possible. The 4” diameter, 5” long waspaloy billet (Figure 5.4(a)) was heated to a steady

state temperature of 1400 °F or 760 °C during the test. Nominal composition of Waspaloy

is given in Table 5.1.

Figure 5.3. Schematics of the test setup to recreate the thermal cycle

106
(a) (b)

Figure 5.4. a) Waspaloy billet mounted to press top bed b) assembled instrumented test

container mounted to the press ram (bottom bed)

Table 5.1. Nominal composition of Waspaloy

107
The test die was made of H13 heat treated to a hardness of 55 Rockwell C. Heat

treatment was done by austenitizing at 1850 °F, followed by gas quenching and 2 cycles of

tempering at 1100 °F for 2 hours each. After heat treatment, these samples were wire

EDMed into 8 sections, ground to flatness and polished so that the samples are all at the

same height when assembled in a container. Photograph of one such sample and assembled

container are shown in Figure 5.4(b) and 5.5.

Figure 5.5. Eight H13 die samples assembled in a container

Assembled bottom die and the billet were mounted on an Instron 1322 material

tester that is capable of a maximum load of 25 tons and a maximum speed of 2” per

second. The billet was mounted such that it is in the hot zone of the inline Applied System

Series 3320 electric furnace. A 0.04” diameter hole was drilled on the test die to within

108
.010” of the surface and a J-type thermocouple inserted. Once the thermocouple was

inserted, thermal cement was applied at the base of the die along the wire length, so that

the wire does not move during the campaign. Again, this was done to ensure quality of

data obtained is acceptable. The entire experimental setup is shown in Figure 5.6.

Figure 5.6. Test setup on Instron 1322

5.3 Thermal Softening Test Campaign

Once the billet reached the set temperature, displacement controlled cycle was

programmed so that the subsurface temperature reaches approximately 400-500 °C. The

programmed velocity profile is shown in Figure 5.7. For the trials, to ensure adequate

109
thermal contact, ram displacements were set so that the loads were approximately 5000

lbs (or a interface pressure of about 4000 psi).

Figure 5.7. Velocity profile used in the tempering test trials, positive indicates motion of

ram towards billet and negative is away from billet

The thermocouple leads were connected to a National Instrument DIN-rail

mounted terminal block that has 14 unconditioned temperature inputs with cold-junction

110
compensation and auto-zeroing. Data was logged using National Instruments Virtual

Bench data logger at maximum allowed logging speed.

Once the cycle was programmed, a campaign was started for a 400 piece run. The

cycle time was approximately 30 seconds. Between each cycle, the dies were air cooled

using pressurized wet shop air. No liquid lubricant spray was used in the test. The thermal

cycle from one of the campaign is shown in Figure 5.8 and Figure 5.9. These two plots

show temperatures recorded at a depth of .010” from the surface. The bulk temperature,

also measured by a J-type thermocouple, was 300 °C and fluctuated very little. It was

noted that the temperature at the surface saw a larger swing in temperature. The swing in

the surface temperatures were much higher (of the order of 50 °C) at the start of the

campaign and dropped to les than 40°C at steady state because of the substrate heating.

Also at the surface, temperature measurements went up and down in a monotonic fashion,

during the cycle.

111
600

500

Temperature (in °C)


400

300 Temperature history

200

100

0
0:00:00 0:32:57 0:58:45 1:24:34 1:50:22 2:16:10 2:41:58
Tim e *in hrs

560

540

520
Tem perature (in °C)

500

480

460 Temperature history

440

420
2:41:58 3:07:46
Tim e

Figure 5.8. Thermal history of 0.010” below surface during contact test campaign

112
520

510

500
Tem perature (in °C)
490

480

470

Temperature history
460

450

440
2:41:58
Time

Figure 5.9. Recorded temperature at .010” below the surface (10 cycles)

The recorded temperature was then matched to numerical model to ensure the

correct heat transfer coefficient is used.

5.4. Finite Element Modeling

The second step in the validation was performing computer simulation of the

process. FEM simulation of the contact test was performed using a 2D explicit finite

element code DEFORM2D v 8.0. Input to the FEM simulation were thermal properties

and flow stress for Waspaloy and H13, interface conditions (heat transfer coefficient) and

process information (billet and ambient temperature, velocity profile (Figure 5.7) and

steady state die temperature). The billet temperature was set at 760 °C and ambient

113
temperature set to the measured 100 °C. Steady state die temperature was set at 250 °C.

It should be noted that the flow stress had no relevance to the simulation as both objects

were considered to be rigid. Apart from the standard forging simulation inputs, starting

hardness (48 HRc) and H13 tempering parameters (C=16.44) were also provided as

inputs. After simulation was run, the maximum temperature and peak to valley difference

in temperature at the surface was obtained. and matched to what was measured.

A series of simulations were performed with different values for the interface heat

transfer coefficient. The maximum temperatures were recorded at .010” below the surface.

The maximum subsurface temperatures for the various heat transfer coefficients are shown

in Table 5.2. Based on the series of simulation, a heat transfer coefficient of 12 was found

to be the most appropriate value, to be used in subsequent modeling.

Heat transfer coefficient Maximum


(N/sec/mm/C) temperature (.010”)
4 454 °C
6 464 °C
8 486 °C
10 494 °C
12 500 °C

Table 5.2. Various heat transfer coefficients used in the model and corresponding

maximum temperatures at a depth of .010” from surface

114
Using this value for the interface heat transfer coefficient, for the given thermal

cycling, another computer simulation was performed to predict the level of tempering seen

at the surface. The algorithm was modified so as provide values for predicted softening

after 400, 800, 1200 and 1600 cycles. Table 5.3 show the hardness loss after 800, 1200

and 1600 cycles at different depths. Figure 5.10, 5.11 and 5.12 show the predicted thermal

softening of the surface layers after 800 pieces, 1200 and 1600 cycles respectively.

Depth of measured location Hardness loss after


from surface (in microns) 800 cycles 1200 cycles 1600 cycles
100 -2.5 -3.9 -4.4
200 -2.4 -2.2 -3.2
300 -2.1 -1.7 -2.8
400 -1.7 -2.0 -2.1
500 -1.2 -1.6 -1.8
600 -1.3 -1.4 -1.5
700 -0.7 -1.3 -1.9
800 -0.4 -1.0 -1.4
1000 -0.2 -0.7 -0.9
1300 -0.1 -0.4 -0.9

Table 5.3. Measured hardness loss at different depths from the surface. Measurement was

made using a microhardness tester under a load of 500 gms

115
Figure 5.10. Predicted hardness loss at the surface after 800 cycles

1 mm

Figure 5.11. Predicted hardness distribution at the surface, after 1200 pieces

116
Figure 5.12. Predicted hardness distribution at the surface, after 1600 pieces

It was noted that the hardness dropped by almost 4-5 Rockwell points after 1600

cycles at a depth of 100 microns.

5.5 Characterization

During the contact test campaign, one of the samples was replaced with a brand

new “sector” every 400 cycles. This approach was used so that one has a moving history

117
of hardness at increasing number of cycles. Figure 5.13 displays the hardness loss in

sample runs of 800, 1200 and 1600 cycles, at different depths from the surface. It should

be noted that at each depths atleast 3 hardness measurements were made. The values in

the table are average values of all hardness values made at the depth. Beyond a depth of 1-

1.2 mm, no distinguishable hardness gradient was observed. No hardness measurements

were made at locations closer than 100 microns because of the influence of free surface on

the hardness measurement. At distance, of at least twice the size of indentation (approx.

80 microns) was maintained above the indented location.

Hardness loss from surface to interior

0
0 200 400 600 800 1000 1200 1400
-0.5

-1
Hardness loss (in Rockwell C)

-1.5

-2

-2.5 HRc Loss (400 cycles)


HRc Loss (1200 cycles)
-3
HRc Loss (1600 cycles)
-3.5

-4

-4.5

-5
Distance from surface (in mm)

Figure 5.13. Measured micro-hardness measurements from surface to the interior

118
Figure 5.14 compares hardness loss predicted by the model and the observed

hardness loss. A fourth degree polynomial was used to get a least square fit of the

measured hardness, which is being compared to the predicted hardness loss at the surface.

The model predicts the hardness loss within .5-.6 HRc.

Hardness loss after 1600 cycles from surface to interior (measured v/s predicted)
Distance from surface (in microns)
0.0
0 200 400 600 800 1000 1200 1400

-1.0
Hardness loss (in Rockwell C)

-2.0

-3.0

-4.0
HRc Loss (1600 cycles)
-5.0 Measured
Poly. Least Square Fit

-6.0

Figure 5.14. Measured value of hardness loss versus predicted hardness loss at different

depths from the contact surface

5.6 Industrial Case Study

Wear model developed was tested out to predict the wear characteristics of a

blocker die used in flashless precision forging of a gear blank. The following section

presents the predicted wear and compares the predicted value to the measured values. The

119
design of tooling and the process along with worn die set was obtained from Eaton

Corporation, Forge Division in Marion, OH.

The modeled process forges a gear blank from AISI 4140, 2” diameter and 4”

long. Parts are being forged on a 1600 ton mechanical press with a total stoke of 14.4”

and a fly wheel speed of 70 spm. Forging temperature was 1010 °C (1850 °F) and the

forging cycle time was 12 seconds. Between successive parts, the dies are lubricated

manually with fine water based colloidal graphite mix. The forging is being done in 3

passes: the first pass pancakes the billet to remove scale and oxide layer that is forms

during heating, the second pass or blocker creates a shape close to the finished part. The

finisher forges part to the final size. Dies used were made of H-13, quenched and double

tempered to a hardness of 48 to 50 HRC (Rockwell C scale). Before operation, these dies

are typically preheated to over 150 °C through the use of flame rings. A schematic

representation of the 3 pass forging of the gear blank is shown in Figure 5.15.

These conditions have been used to model the process to establish a benchmark for

prediction of wear. Figure 5.16(a) show the worn finish and blocker dies and Figure

5.16(b) shows a close-up of the blocker die after a run of 2400 pieces. Although the finish

dimensions are more critical, wear in blocker pass is considerably more. This is because,

blocker pass, in general, involves a substantial movement of metal. The finish pass is used

mainly to ensure the part is finished to the final dimensions. Because of this reason, for

validation of the wear model, the blocker stage forging was used.

120
(a) (b) (c)

(d) (e)

Figure 5.15. Schematics of a) buster, b) blocker c) finisher d) end of stroke and e)gear

blank forging (courtesy: Eaton Corp., South bend, Indiana)

121
(a)

(b)

Figure 5.16. (a) Worn blocker and finish dies (b) Close-up of blocker used in validation

122
Validation of the wear model involves computer modeling of and wear and

correlation of the predicted wear to the actual wear profile. Properties that are critical to

the use of prediction were the tempering parameter of H13, initial hardness of the dies and

the hot hardness of H13. A tempering parameter of 16.44 was used for the model. Starting

hardness of the dies ranged from 48-50 (per Eaton Corp). The hot hardness of H13 was

obtained from standard data sheet for H13 and sown in Table 5.4.

Table 5.4. Hot hardness of H13

Process information used in the computer simulation corresponded to what was

used in the actual forging. Due to lack of data, a standard heat transfer coefficient of 11

KW/m2 °C was used. This value is generally accepted as a standard heat transfer

coefficient where water based graphite is used. Figures 5.17, 5.18 and 5.19 show the wear

profile in the top blocker die predicted by the model.

123
Figure 5.17. Predicted total wear profile in blocker die for the gear blank forging case

study

124
Figure 5.18. Predicted total wear distribution from center to outside diameter, in blocker

die

125
Figure 5.19. Corner wear predicted using modified Archard model

Simultaneously, profile of the worn blocker die was measured using a Sheffield

Cordax coordinate measuring machine (CMM) with a 1 mm ruby tip (Figure 5.20). The

CMM has a rated linear accuracy of 0.0003”, a volumetric accuracy of 0.00043” and a

repeatability of 0.00012”. . To ensure consistency of data, measurements were made in 4

different radial directions. This also ensures that the effect of misalignment or mismatch

during forging (which is not part of the 2D finite element model) on wear gets smoothed

out, to some extent.

126
Figure 5.20. Worn die and Sheffield Cordax contact type CMM used to measure the

profile of the worn die

Once data was collected, the data was converted to a CAD format by converting

this data to a IGES file and importing this file into a CAD software Solidedge. This

process allowed comparison of the worn profile to the original CAD file. To be consistent,

the wear values reported was measured perpendicular to the surface of the die at the

measured location. It was assumed that the CAD profile used to evaluate the worn die was

a true representation of the shape of the original die. Figure 5.21 shows the graphical

approach used to measure wear.

127
Measured

wear

Figure 5.21. CAD approach used to compare the predicted and measured wear profile

Table 5.5 shows the results of the wear measured. The table also shows the values

obtained for the predicted wear in the blocker stage die. The predicted values have been

normalized so that the peak values of the predicted wear and the measured wear are the

128
same. This method, hence, only tests the validity of the wear model in predicting the

relative wear at different locations. However, it should also be noted that the absolute

values of the wear predicted can be compared across other similar forgings, with

reasonable confidence.

X (mm) Predicted Normalized Measured


5.34 293000 0.095 0.02
10.69 636000 0.205 0.14
17.37 1040000 0.336 0.28
24.05 1410000 0.455 0.44
26.73 1580000 0.510 0.5
30.74 2040000 0.659 0.68
33.38 2380000 0.768 0.75
34.65 2470000 0.797 0.82
35.86 2540000 0.820 0.75
37.01 2390000 0.772 0.69
38.08 2240000 0.723 0.6
39.06 1770000 0.571 0.57
40.69 975000 0.315 0.38
41.24 516000 0.167 0.3
41.85 196000 0.063 0.21
42.23 69900 0.023 0.1
42.49 43300 0.014 0
43.33 0 0.000 0

Table 5.5. Predicted, normalized and measured values of the wear in blocker die

129
Wear Profile - Measured vs Predicted

0.900

0.800
Relative Wear (norm alized to peak)

0.700

0.600

0.500
Normalized
0.400
Measured
0.300

0.200

0.100

0.000
0 10 20 30 40 50
-0.100
Distance from cente r (in m m)

Figure 5.22. Measured and predicted wear profile in blocker die of hot precision gear

blank forging case study

Maximum wear is seen at the punch nose radius. Predicted wear and measured

(Figure 5.22) values show a sharp drop beyond the midpoint of the radius. This is due to

the fact that outside the center of the nose radius, the pressures are lower. It was noted

that the model predicts the peak of the worn profile within 1 mm in the radial direction.

However, the model over predicts the wear near the center by a significant amount. The

plot only extends till the base of the central hub. Both the measured and predicted wear

beyond this region is negligible, as these regions are subjected to minimal sliding and

considerably lower normal pressures and temperatures compared to the center of the

punch.

130
CHAPTER 6

INTEGRATED SYSTEM: FRAMEWORK AND APPLICATIONS

6.1 Summary

In this dissertation, an integrated approach to prediction of failure of forging tools

experiencing high temperatures, pressures and sliding is proposed, implemented and

verified. This work was done in two phases.

In the first phase, tempering parameters were used to predict the intermittent

tempering and loss of hardness of forging dies made of H13 tool steel. This model was

incorporated into DEFORM-2D finite element software as a module. Although the

program uses known values of tempering parameter published in the literature, an

approach to evaluating tempering parameter from data sheets has been outlined. The

approach was used to demonstrate the effect of various forging process parameters

(equipment selected, lubricant dilution ratio and cycle times) on the hardness loss at the

surface. These results clearly shows the varying needs of tools used in hydraulic press

forgings, as against a mechanical press.

131
A simple test that imposed a thermal cycle on the dies, which is commonly seen in

hot forging, was devised, designed and implemented. A controlled instrumented

experiment was performed and tools characterized to obtain hardness degradation. The

test was also modeled using DEFORM-2D, to predict the loss in hardness of the

temperature. The predicted results have been verified with the hardness loss measured in

the test specimen. Results show a reasonably good trend and accuracy of prediction.

The second phase of the program incorporated the results of the first phase, and

known relationships between the hardness and temperatures available in the literature, to

modify Archard’s existing model for prediction of wear. This module incorporates

softening behavior of the substrate and hardness loss associated with high temperatures to

predict the wear in forging dies. This approach has been verified on an industrial precision

gear blank forging application, with mixed results.

It is the belief of the author that, once calibrated for die material of interest, this

approach can be successfully used to predict life of forging tools for a given process

condition. Also, one can study the effect of die design (flash and gutter design, buster and

blocker geometry) as well as process selection (forging temperature, lubricant, equipment,

cycle times etc.) on the life of dies. Based on the model and program developed, one can

then modify the process or tool material to reduce forging equipment downtime. Also,

through development and/or application of new die materials, one can potentially forge

new product lines like fine pitched gears, that are currently technically challenging.

132
6.2 Future work

Preliminary validation performed on the industrial case study, for H13 die steel,

shows good promise. Results obtained are applicable to cases where there is published

information on physical and thermal properties (thermal conductivity, specific heat, CTE)

as well as phenomenological properties like flow stress and heat transfer.

But the next generation of forging dies for extended die lives probably will not use

conventional die materials. Approaches that the author foresees being used in the near

future, involve the use of advanced die materials like nickel based superalloys, stellites and

tungsten carbides as coatings or cladding. There is a dearth of literature on the thermal,

physical and phenomenological properties for these materials. Extensive work needs to be

done on both methods to apply these advanced materials on cheaper substrates as well as

characterizing the properties of the engineered surfaces. Functionally gradient material

(FGM) approaches involving these high performance die materials on relatively cheaper

substrates need to be developed and tested. Preliminary work on one of the newer

methods of application of these on conventional substrates, Laser Engineered Net Shaping

(LENS) is given in the Appendix.

To characterize the heat transfer coefficient, thermal conductivity and the specific

heats, a test setup similar to the one shown in Figure 4.20 can be used. However, because

there are more variables involved there needs to be more instrumented data available, to

inversely obtain the relevant thermal properties.

133
Also, once these coatings are deposited, the wear resistance of the surface needs to

be characterized and documented, so that these can be used by forgers with reduced risk.

A test setup is shown below (Figure 6.1), that has been designed to simultaneously

characterize wear resistance of several surfaces under controlled condition. This test dies

have been designed to run on a 350 ton hydraulic press (currently being installed at the

Ohio State University).

Figure 6.1. Proposed wear tests to simultaneously characterize several surfaces in a single

test run

6.3 Conclusions

Optimization of the forging process through the use of process modeling and

advanced materials will be a key for reducing costs associated with equipment down time,

134
die changes and quality. Selective incorporating functionally gradient materials to improve

performance of forging tools will also enable forging new class of products like fine tooth

splines and gears, which will immensely improve the competitiveness of forgers. US has

continued to lose hot forging to cheaper global competition, because these have now been

thought of as commodity products. The reverse this trend, changes will have to happen to

improve the value of the forging process. This dissertation outlines some of the effort

done at the Center for Excellence in Forging Technologies, towards these objectives.

135
APPENDIX A

DEVELOPMENT OF LENS PROCESS MODELS

A.1 Introduction TO LENS

To understanding the impact of various factors on the properties of the coating, a

process model based on underlying physics is necessary. LENS based deposition process

can be classified into 4 stages:

• Stage 1. Powder travel from the nozzle to the substrate surface: During this time

period, the powder is heated continuously.

• Stage 2. Melting of the substrate: Incident laser heats the surface and melts the skin.

There is a possibility of plasma formation that will affect the flow of powder.

• Stage 3. Mixing of semi-heated powder: During this stage, solid or semi-solid powder

material splashes on the heated substrate and mixes with the substrate forming a melt

pool. The melt pool is subjected to further heating by the laser beam. In some cases,

exothermic energy may be released by the powder that will also heat the pool.

• Stage 4. Solidification of the melt pool after it has left the laser zone: The last phase

occurs when the table moves away from the beam and energy flow into melt pool

136
stops. Melt pool created in the previous stage solidifies rapidly in the wake of the laser

beam path forming a dense adherent layer on top of the substrate.

To understand the effects the control variables will have on the properties of the

ensuing coating, underlying physics of the formation needs to be understood. The primary

physical phenomena involved in the laser deposition process are

• Heat transfer to powder

• Trajectory

• Plasma formation

• Laser heating of substrate

• Impact of heated powder on substrate and splat formation

• Chemical energy in mixing

• Physical mixing of semi-solid powder and substrate (convection)

Heat transfer to substrate mass - Solidification

A.2 Prior Modeling Efforts

Several researchers have either modeled Laser Engineered Net Shaping process or

similar processes or have used modeling techniques that can be used in modeling LENS

process. These involve the field of physics of materials, thermal analysis, fluid flow and

material science. The following section sums up the various work found in the literature in

modeling of the 4 stages.

137
A.2.1 Stage 1: Powder free-fall in laser zone

During this stage the particles fall through the laser focal area and are subjected to heating.

Heat infused into the powder depends on the time it spends in the laser zone, laser power

and the heat absorption. Also, it depends on the size and shape of the powder. The time

powder spends in the laser focus depends on the location of the nozzle with respect to the

Effect of control variables on powder pretreatment

Several researchers have attempted to predict the extent deposited powder heats

when it is falling in the laser zone. LENS parameters that influence the powder

temperature and distribution are

• Location and geometry of the nozzle

• Flow rate of the powder

• Heat absorptivity of the powder

• Size distribution of the powder

• Laser power, beam diameter and focus

Argon flow rate

Time the powder spends in the laser zone depends on the powder velocity,

distance and orientation of the nozzle, size of the powder, back pressure due to plasma

and recoil of shroud gas. The heat absorption depends on the absorptivity (a function of

laser wavelength and temperature), powder surface to volume ratio. The absorptivity of

the powder is temperature dependent and rises with increasing temperature. It also

138
increases with lower wavelengths and hence the higher efficiencies in use Nd-Yag as

against Co2 laser.

Fisher and others [Fischer, et al., 2002, Fischer, et al., 2003] have performed

numerical simulations in sintering of Ti powder using SLS process factoring in absorption

of radiation by metallic powder, plasma formation (function of average laser power,

pulsing rate, beam radius) and resulting recoil pressure, and the thermal properties of

powder and the binder medium. Numerical simulations were done develop process maps

that would help in establishing laser pulse frequency and laser power. Approximate

physics-based and empirical-models have been used for this purpose, to relate the laser

setting and powder size to temperature distribution in the substrate, plasma recoil pressure

and subsequent surface profile. The results have been validated using sintering

experiments on Ti powder performed using a 100W Nd-Yag pulsing laser with reasonable

results.

Equations used in the model are

P0
vp =
τ p πrb2 I p
(A.1)

P0 is the average power, τ p is the pulse duration, rb is the beam radius and I p is the

plasma threshold beyond which plasma can be expected,

I p = 4 x10 4 / τ
(A.2)

139
Prec = .54 p (Ts ) (A.3)

where P is the recoil pressure and p(Ts ) is the vapor pressure at temperature Ts

Maximum temperature at the surface of a powder is approximated by

2 AI 0 k thτ p
∆Tmax =
k π (A.4)

where A is the absorptivity, k is the thermal conductivity, kth is the heat diffusivity and I 0

is the intensity defined by

P0
I0 =
πrb v mτ p
2
(A.5)

Yevko and others from University of Toronto [Yevko, et al., 1998] have modeled

the temperature field in SLS process using FEM (IDEAS ) to solve generalized heat

transfer problem with prescribed boundary conditions. IDEAS allows laser power source

to be modeled as a heat flux that is a function of time. Yevko defines heat load as a

function of laser power and beam width and varies the load based on the scan speed. In

this exercise, the density and specific heat of the powder were calculated based on law of

mixture using properties of stainless steel and argon. The thermal conductivity was

calculated based on the assumption that all the powder particles are spherical and

compacted (Churchill’s work).

140

K =1+
1.306φ 10 / 3
1−φ − − .07296φ 14 / 3
1 − .4072φ 7 /3
(A.6)

where K is the conductivity, and φ is defined as

kd
−1
kc
φ=
kd
+2
kc x volume fraction of dispersed phase (A.7)

where kc is the thermal conductivity of continuous argon phase k d is the thermal

conductivity of the dispersed phase.

Yevko’s thermal simulations yielded temperature profiles that were used to

calculate the clad width. The clad height was calculated based on increase in density

because of associated melting and densification. It should be notes that this analysis

assumes no plasma formation and that the molten powder is spherical in shape. The results

of simulation was validated using measurements of weld shape performed using a 1kW

Nd-YAG pulsing laser.

Grujicic and others [Miller, et al., 2001] have come up with a simple model for

predicting the temperature of a single powder particle propelled through a laser focal

zone. Using simple geometric assumptions, they calculate the time a particle spends in the

laser t r as

t r = 2w0 / v p sin(θ )
(A.8)

141
where w0 is the radius of the laser beam, v p is the particle velocity and θ is the angle

between the trajectory and the laser incident angle. The energy absorbed E is

rp
Ep =α( ) 2 Pl t r
w0 (A.9)

where α is the energy absorption parameter, rp is the particle radius and Pl is the laser

power. The authors calculate the total energy required for melting of the powder by

summing the energy required to raise the powder temperature to melting temperature and

the energy needed to completely melt it. By incorporating empirical results by

Keicher[Grujicic, et al., 2001], authors were able to constrain the problem so that the

substrate did not melt. Keicher’s study suggests that for Inconel 625, substrate melting

does not take place if the laser irradiance is less than 1 W/m2. This can be represented in

equation form as

C l Pl
I 1 = w02 (A.10)

This equation is applicable if the laser beam is focused at the surface. If it is

focused at a distance h 0 above the surface,

C l Pl
I 1 = w0 + h 0 tan( β / 2) 2 (A.11)

The problem was solved as an optimization problem using MATLAB to maximize

the powder size for complete in-flight melting.

142
Fu and others[Fu, et al., 2002] have attempted to calculate the laser beam power

attenuation due to absorption of laser photons by the powder particles before it reaches

the surface. Their model explains loss of power by laser by taking into account the time

the powder particles spend in the laser beam focus, its size and number, its absorptivity

and the powder delivery angle. They have focused their efforts on the temperature

distribution of the powder particles because they believe that the primary driving force in

pool motion in laser cladding is the temperature gradients in the powder fed.

Kathuria and others [Kathuria, 1997] found that the interaction time of laser

(Figure A1) with the powder along with the cooling rates is critical in determining the

microstructure of the clad layer. They found that the longer the interaction time, the

mushier the microstructure is.

Figure A.1. Microstructure v/s processing conditions for laser cladding (D = beam

diameter in mm, V is the table speed in mm/s)[Kathuria, 1997]

143
A.2.1. Stage 2: Melting of Substrate and bead formation

Typically, in LENS, the laser beam is focused at the surface. A fraction of the

radiant energy goes into melting of the surface and the remaining energy is used to preheat

the powder that falls through the focal region. It is believed that almost no melting of

powder occurs during this phase. The incident laser beam creates a molten pool of metal

on the substrate which captures the feed powder. The pool width could range anywhere

from ½ time to 5 times the beam diameter[Lewis and Schlienger, 2000]. The pool

geometry and pool superheat dictate the efficiency of the particle catchment, build rate and

the bead properties. Overheating increases dilution but flattens the bead by reducing the

surface tension.

Lewis and Schlienger [Lewis and Schlienger, 2000] indicate that the width of the

weld pool created by the laser beam is primarily dependent on the surface tension of the

molten material. Also, they indicate that the higher the heat sink is, the smaller is the melt

pool width. They indicate that increasing laser power, lowering traverse speed and using

thinner layers help achieve better finish.

With substrate melting as the primary variable, Grujicic [Miller, et al., 2001] and

others have solved a second order heat transfer differential equation to determine the

extend of the substrate melting. In their analysis, they have considered the radiative and

convective heat loss from the substrate surface and heat flux due to the laser beam, which

144
they have defined as a function of the beam diameter, laser velocity and the power levels.

They have defined the peak laser intensity I 0 of a gaussian beam as

4 Pl
I0 =
2πw0 (A.12)

where P is the laser power level and w is the laser beam width.

Gaussian beam power level P(r) at any point r from the center beam is given by

P0  r2 
Pr = exp − 2 
πw0 2  w0  (A.13)

For simplicity, molten material is not explicitly considered in their analysis. They

solved the thermal conservation equation through finite difference approach to find size of

the melt pool.

Vasinonta and others [Vasinonta, et al., 2000] used ABAQUS to model the

thermal distribution assuming a point source approximation. They also neglected the effect

of convection to air and convection within the melt pool, based on findings of Dobranich

and others. They used non-dimensionalized variables (Rosenthal) to model and predict the

temperature distribution and develop process maps for LENS. Using their model they

predicted melt pool length and compared these to actual lengths that were measured on a

LENS system. Their model predicted reasonably well, expected increase in melt pool

lengths with increasing power and decreasing velocity. Their model also suggests that

substrate preheating has a pronounced effect in reducing residual stress.

145
Ki and Mazumder [Ki, et al., 2001] have modeled and solved the convective heat

and mass transfer that occurs in laser material removal process using level set method, to

compute the liquid-vapor interface. Their element based mass balance approach accounts

for evaporation, melting of substrate and convective flow in the liquid phase. Heat balance

in the element is achieved by balancing radiant and convective heat transfer, laser

radiation, heat loss in vaporization and heat lost to substrate by conduction. It should be

noted that their model is applicable to non-additive processes like laser melting or laser

machining. Adapting their model to LENS creates another dimension when one attempts

to incorporate powder heating, heat lost to powder and splatter.

Although Fu and others [Fu, et al., 2002] have not modeled substrate heating, they

have modeled powder heating. The attenuated laser beam power density given by their

calculations can now be used to model how the laser beam heats the substrate. To

calculate powder catchment efficiency and clad height in a V-groove of a edge joint, Lin

and others [Lin and Hwang, 2001] have performed simulations of the powder stream

without laser radiation using fluid dynamic computational software, FLUENT which is

based on a specific control-volume approach. Cladding experiments were performed for

mild steel substrates with thickness of 2 and 6 mm under 1 kW CO2 laser irradiation for

304L stainless steel powder. They have decoupled flow and heat transfer to first obtain a

powder flow distribution which was then transformed to clad height using heat transfer

equations. Qian and others [Qian, et al., 1997] have studied the effect of the translation

speed of the laser beam and the powder feed rate on the physical and microstructural

146
qualities of the clad layer. This study attempts to experimentally arrive at optimal cladding

parameters for maximum hardness.

Li and others [Li and Ma, 1997] studied surface finish of blown powder cladding

method theoretically and experimentally. Their approach to modeling surface finish was

primarily geometry-based. Although the model does not help in understanding the effect of

process parameters on surface finish, it gives us a general direction to any approach to

improving the texture of clad surfaces. They found that the surface roughness (turbulence)

of an overlapped cladding layer decreased with the increase of the overlapping ratio in an

oscillating manner (Figure A3). They have modeled each clad layer as a parabolic

approximation.

Figure A.2 Effect of travel speed on aspect ratio of the clad bead [Qian, et al., 1997]

147
Figure A.3. Variation of surface finish in powder-based cladding[Li and Ma, 1997]

Several empirical relationships have been suggested that relates the bead height

and width to the process variables. Boddu and others[Boddu, et al., 2001] refer to work

by Steen and others that relate bead width W to traverse speed v and laser sport size D.

W = D(1 − aν ) (A.14)

where a is a constant. Hu and others [Hu, et al., 1998] have also showed that the clad

geometry can be approximated by a parabola and that the clad height H(t) can be

approximated by

148
H = 3 A 2W
(A.15)

where A is the cross section area of the clad pass and W is the clad width. Hu and others

performed several experiments that showed that the clad height increases with increasing

powder flow rate, increasing power density or with decreasing traverse speeds.

The dilution of the clad layer into the substrate depends on several factors: the

thermal conductivity of the substrate material, the initial temperature of the substrate, the

reflectivity of the material, the powder flow rate, the interaction time of the powder in the

beam and laser power [Sexton, et al., 2002]. Sexton reports that, for low laser power

values, no fusion of the substrate occurred and the coating did not adhere to the substrate,

while for larger power values the substrate melted increasingly as the power levels

increased causing increased dilution.

Ming and others[Ming, et al., 1998] have experimentally verified the effects of

laser travel speed on dilution and bead aspect ratio predicted by other researchers. They

found that the aspect ratio (width to height) increased with increasing travel speed (Figure

A4) and powder feed rates. Also, the dilution increased with increasing speeds and

decreasing powder feed rates. It should be noted that the less the dilution, greater is the

wear and oxidation resistance of the coating.

149
Figure A.4. Effects of laser traverse speed on aspect ratio of bead and dilution [Ming, et

al., 1998]

A.2.3. Stage 3: Mixing of Substrate and Powder

The third stage of the process is mixing of the substrate and the powder captured.

The laser beam passing over the substrate or the prior deposit and melts a small region.

The size and the superheat of the melt pool controls how effectively the powder fed is

captured on to the surface. Higher the surface area, higher is the expected powder capture

efficiency. Also, higher the superheat, higher is the ability to melt the powder particles and

homogenize the weld matrix.

Mahrle and Schmidt [Mahrle and Schmidt, 2002] recognized temperature as the

key parameter in welding in affecting the weld properties and have modeled the laser weld

process using finite element approach. They have used steady state models and non-

150
dimensional numbers (Prandl, Peclet and Reynolds numbers) to predict flow induced heat

transfer, melt temperature, velocities and weld bead shapes.

In some cases, enthalpy plays a big role in the thermal behavior and the ensuing

structure. Dupont and others have worked on elemental blends of Nickel Aluminide.

However, control of the thermal phenomena, the microstructure and porosity becomes

more difficult because of the effect of the additional variable (the enthalpy of mixing).

Liu and Dupont [Liu and Huang, 2003] investigated in-situ layering of Nickel

Aluminide using Nickel and Aluminum powder and Ti-6-4 using elemental blends. Their

study revealed that NiAl produced by this method was rather brittle and exhibited porosity

that was attributable to the material as well as the high thermal stresses during the process.

Similar work has been done by Fraser and others [Collins, et al., 2003], [Banerjee,

et al., 2002, Collins, et al., 2003] that focused primarily on the reaction kinetics in creating

Ti and Mo alloys (like Ti-6Al-4V) both from elemental blends as well as premixed alloys.

Their study has revealed that it is possible to create alloys in-situ, however, control of

compositions is extremely challenging.

Kar and Mazumder [Kar and Mazumder, 1996] have analyzed the shape and

morphology, of Laser CVD process and have presented a model that allows one to select

the appropriate process parameters to obtain a good quality film. They have found that an

optimum condition is found to exist for depositing thin films by using the LCVD

technique.

151
A.2.4. Stage 4: Solidification

The last phase of deposition process is solidification. In this phase, the molten

substrate-powder mixture cools to form a rough layer. Modeling of the last phase involves

heat transfer, microstructure change and diffusion.

Researchers at Sandia Labs [Griffith, et al., 2000, J. A. Brooks, et al., 1999,

Schlienger, et al., 1999] have tried to model the thermal effects of the LENS process on

the microstructure of H-13 through application of the thermal histories extracted from

experiments and phenomenological models of hardening process. They have shown that

microstructure of the weld area as well as secondary hardening of the substrate close to

the weld can be controlled through the use of appropriate table speeds, power levels and

hatch spacing. They found that the peak temperature is the most critical process variable

that controls the hardness of the substrate.

A.3 Numerical Model Development

For success of coatings and surface engineered dies, properties that are most

critical are the wear, oxidation and thermal fatigue resistance of the clad layer, the

adhesion of the clad layer to the substrate and the geometrical integrity of the weld. Local

temperatures, weld pool shape, depth of melting and the powder feed rates have the

biggest impact on these parameters. Process models created need to address the physical

phenomena that affect these. Also, model created should have a way to relate these

152
variables to the control variables (laser power, table speed, laser path, powder size and

laser beam diameter), so that the tool developed can be used for design of the process.

One approach to modeling these phenomena is through finite element codes. FEM

is a robust approach that allows one to model non-linear processes more accurately, taking

into account irregular geometries, nonlinear material properties and coupled phenomena.

Also, FEM allows one to extract transient thermal and flow behavior, keys to predicting

weld properties.

For this application, it has been well documented that flow within the melt pool has

only a minor effect on the process variables that are of interest. With this assumption, a

numerical model of the LENS process has been developed using finite element code

DEFORM-3D. For modeling purposes, the physics of the process has been split into 2

phases:

• Heating and melting of the substrate by a laser beam

• Addition of heated powder to the melted weld pool

The first stage of the process is modeled as a heat transfer problem (Figure A5).

The power and laser beam absorption were combined into a single simulation control

variable called effective power (given by equation A.16). Effective power is defined as the

product of the laser power incident on the substrate and the effective absorption of the

substrate (A). Effective absorption, as defined above, is not only a function of the surface

153
condition, but also a function of the power absorbed by the powder as it drops through the

laser focus.

Convection+

Conduction

Figure A.5. Laser beam representation and thermal boundary condition modeled in FEM

Heating of the substrate by the laser beam has been modeled using a heat transfer

window that is exactly equal to the size of the laser beam.

Pe = A.Pl (A.16)

A preliminary simulation was conducted assuming a rectangular substrate of size

2”x2”x1”. The simulation revealed that the temperatures outside a region of

approximately 1” were almost room temperature. Based on the temperature profile

obtained from the simulation run, subsequent runs were modeled based on a smaller

154
specimen geometry This allows us to refine the mesh size even more for the same

computation times. Also, for reducing computation, one half of the substrate has been

modeled.

To develop the model, the following assumptions were made:

• Although the beam power intensity is gaussian in nature, it has been modeled as a

beam with uniform power intensity.

• The melt pool and the unmelted substrate is being modeled as a single phase. Thermo-

physical properties of melt pool (thermal conductivity, specific heat etc.) has been

incorporated in to the model as a temperature depended property

Effects of convection induced by surface tension, non-uniform densities etc. on local

temperature distribution is neglected

A.4 Parametric Numerical Study of LENS Process

To study the effect of the different process variables on the characteristics of the

weld bead (width and depth of the weld pool, superheat), several numerical simulations

were completed using a 3D explicit FEM code, DEFORM-3D.

The process of LENS was simulated as a purely thermal phenomenon, to study the

effect of the process variables on the bead width, bead depth and the post process cooling

rates. Process variables that have reportedly large effect on these are the laser power, the

powder absorption, table speed, the laser focus/beam diameter and the preheat of the

155
substrate. These have a large effect on the dilution in the weld pool as well as the

microstructure of the weldment. These 2 variables are the most critical aspects of the

weldment that affects its thermal and mechanical properties. Microstructure with finer

grain size results in a stronger structure. Low dilution in the weld pool will ensure the

properties of the clad layer are close to the engineered composition. The process variables

that were varied in the simulation were the absorbed power, the table speed and the laser

beam diameter. Simulation matrix used in this study is shown in Table A.1.

ID Elements Nodes V (mm/sec) H# Beam Dia. Max h


1 55772 12182 12.5 1 1 mm 2109
2 55772 12182 12.5 2 1 mm 1054
3 55772 12182 12.5 3 1 mm 527
4 55772 12182 12.5 4 1 mm 791
5 55772 12182 4 3 1 mm 527*
6 55772 12182 8 3 1 mm 527
7 55772 12182 16 3 1 mm 527

8 55772 12182 20 3 1 mm 527


9 57614 12419 12.5 3 .4 mm 527*
10 55772 12182 12.5 5 .6 mm 1465
11 55772 12182 12.5 3 .8 mm 527
12 55772 12182 12.5 2 .7 mm 1054

Table A.1. Simulation matrix used for modeling LENS process in DEFORM-3D

156
It should be noted that the absorbed power being modeled encompasses several

phenomena: laser power, absorptions, powder size and emissivity of the powder and the

time powder spends under the laser beam.

Simulation ID Melt Melt Melt Depth Max Temp ° C


width Length
1 1.5704 2.9058 .746 5910*
2 1.3807 1.7864 .5209 4780*
3 .98063 1.0547 .2369 2550
4 1.1831 1.4315 .4044 3690
5 2290
6 1.0763 1.1358 .29517 2720
7 .9441 1.0343 .20493 2400
8 .8355 .915 .1631 2270
9 Did not melt substrate 1180
10 .7869 .8783 .2714 4220
11 .6478 .6560 .12545 2130
12 .823 .9138 .254 3620
* Not applicable, outside of property domain

Table A.2. Simulated maximum melt pool superheat temperatures and melt pool sizes

Figure A.6. show typical plot obtained from the modeling run, showing the

isotherms in the substrate. Assuming a liquidus temperature of 1470 °C, melt pool

157
dimensions were calculated. Figure A.7 show the variation of temperature with time. Peak

temperatures are seen at locations directly under the laser beam.

Figure A.6. Typical thermal profile obtained from the simulation, used to obtain the

length, width and height of the melt pool

158
Also, it was noted that the temperature at locations behind the laser beam was

substantially higher than those ahead of the beam. Cooling rates at the wake of the melt

pool was of the order of 1000-1500 °C/sec.

Figure A.6. Typical thermal profile obtained from DEFORM-3D simulations of LENS

process (temperature in °C)

Simulations were performed at boundary conditions outlined above and the

dimensions of weld documented. Care was taken so that the element size was atleast 1/4th

of the laser beam diameter. The maximum temperature and the bead size were then plotted

with respect to absorbed power, table speed and laser beam diameter. It was seen that the

melt pool temperature, length of the melt pool and the depth increases with the power

159
absorbed. Of the three parameters, the melt pool depth was least sensitive to the power, as

indicated by the lower slope of the curve (Figure A.7).

6000 2
Temperature
1.8
5000 Length
Depth 1.6

Dimensions (in mm)


Temperature (in °C)
1.4
4000
1.2
3000 1
0.8
2000
0.6
0.4
1000
0.2
0 0
120 180 240
Power (in watts)

Figure A.7. Simulated effect of absorbed power on weld superheat (table speed = 12.5

mm / sec, preheat temperature = 20 °C, H13 substrate, laser beam diameter = 1 mm)

All three plotted weld properties decreased with table speed (Figure A8). When

the laser beam diameter was increased, for the same power, the superheat dropped.

However, the dimensions of the weld bead increased. These trends confirm what has been

reported in the literature. However, these need to be validated.

Temperatures observed in these simulations are too high, indicating model

deficiencies that do not take into account power lost to atmosphere, and power lost in

heating the argon stream. Also, it may indicate that latent heat of melting needs to be

modeled more accurately.

160
2800 1.2
2700
1

Dimensions (in mm)


2600

Temperature (in °C)


0.8
2500 Temperature
2400 Length 0.6
2300 Depth
0.4
2200
0.2
2100
2000 0
8 13 16 20
Table Speed (in mm/sec)

Figure A.8. Simulated effect of table speed on melt pool superheat (effective power = 120

watts, preheat temperature = 20 °C, H13 substrate, laser beam diameter = 1 mm)

4500 1.2
4000
1
Temperature (in °C)

3500
3000 0.8 Dimension (in mm)
Temperature
2500
0.6 Length
2000
Depth
1500 0.4
1000
0.2
500
0 0
1 0.707 0.6
Beam diameter (in mm)

Figure A.9. Simulated effect of beam diameter on melt pool superheat (effective absorbed

power=120 watts, table speed = 12.5 mm/sec, preheat temperature =20 °C, H13

substrate)

161
A.5 Validation - Experimental Procedure

There are four primary components of the LENS™ assembly: the laser system, the

powder delivery system, the controlled environment glove box, and the motion control

system. A 760 W Nd:YAG laser, which produced near-infrared laser radiation at a

wavelength of 1.064 mm, was used for all the depositions. The energy density used was in

the range of 30,000 to 100,000 W/cm2.

The basic LENS system 750 used in the tests below consists of a high power

Nd:YAG laser, a 3-axis computer controlled positioning system and multiple powder feed

units. The positioning stages are mounted inside an argon-filled glove box (nominal

oxygen level of 2-3 ppm), while the laser beam enters the glove box through a top

mounted window. A powder delivery nozzle is used to inject a metal powder stream

directly into the focused laser beam. The lens and powder delivery nozzle move as an

integral unit in the z-axis, while the part, positioned under the laser beam, is transitioned in

x and y.

Specifications of the LENS 750 system are as follows:

• Max. build Size: 300mm by 300mm by 300mm (12”x12”x12”)

• Preferred powder size: 75 microns ( - 100 / +325 mesh)

• Hatch width: .015”, build layer thickness: .010”

• Laser power: 270-370 W (1.064 MHz)

• Laser speed 10-50” / min

162
• Powder feed units: 2 buckets

The various components of the LENS system (Figure A.10) are as follows:

• LENS 750 Glove Box and Dri-Train: Chamber where the deposition occurs. It

includes the Nd:YAG laser, optics, the motion controlled stage, the anti-chamber, the

powder feeders, and the atmosphere control systems

• LENS Workstation Control Console: LENS Workstation Control Console is the

interactive computer system for the control of the CAD, Slice, and DMC files, the

motion of the stage, and the atmosphere control of the Glove Box, Dri-Train, and

Ante-Chamber.

• US Laser Power Supply: The system is powered by US Laser Power Supply that

supplies and controls the power to the Nd:YAG laser on the 750 Glove Box and Dri-

Train assembly. The system also supplies cooling water to the laser and the laser

shutters. The laser used in this LENS system is a YAG-Nd solid-state type. These

lasers have a yttrium-aluminum-garnet crystal (YAG) doped with neodymium (Nd).

Instead of the usual 'flash-tube pumping', this new laser uses small semiconductor

diode lasers to excite the neodymium. This allows much greater efficiency - thus

smaller and more powerful lasers than before needed to melt

163
Figure A.10. LENS system a) schematic and b. experimental facility at the Ohio State

University [Banerjee, et al., 2002]

The table speed, the laser power and the hatch spacing the three primary controls

affecting the morphology of the coating and the surface finish. Hence a series of coating

trials were performed choosing 2 levels for each factor. It was felt that, increasing the

164
powder feed will have same effect as reducing the travel speed. Hence it was not factored

into the first series of experiments.

Several H-13 samples having identical geometry were machined. The samples were

heat treated to 46-48 HRC. The head geometry is shown below. To eliminate surface

condition from causing any variability, all samples were polished to similar roughness (~

0.1 microns). A fixture was also fabricated to securely hold the samples at predetermined

location. The fixture designed had several “pigeon-hole” type features that allowed the

pins top be securely located on to the surface. The plate itself was mounted onto the

LENS glove box for processing.

One of the powder reservoirs was loaded with Stellite’s Nistelle N13Al (a nickel

aluminide grade powder). Powder used was rated at -100/+270 Mesh. The substrate was

not preheated. Background oxygen level was kept at 3-4 ppm and the oxygen level during

deposition process and the Argon flow rate was kept at 6-8 l/min. Programmed layer

thickness was set at 0.008”. Powder density was measured to be approximately 7.3

gms/cc.

165
Hatch
Laser speed Current Powder Power Beam
# Spaci Ra (µm)
(mm/min) (amps) flow rate (watts) diameter
ng
1 875 34 .010” 2.5 rpm .1 310 1 mm
2 875 34 .010” 2.5 rpm .1 310 1 mm
3 875 38 .010” 2.5 rpm 0 380 1 mm
4 500 34 .010” 2.5 rpm .1 310 1 mm
5 500 38 .010” 2.5 rpm .1 380 1 mm
6 500 34 .015” 2.5 rpm .1 310 1 mm
7 500 38 .015” 2.5 rpm 0 380 1 mm
8 875 34 .015” 2.5 rpm 0 310 1 mm
9 875 38 .015 2.5 rpm .1 380 1 mm

Table A.3. Measured surface roughness and profile of LENS coated NiAl samples,

powder size used was 100-270 ASTM mesh

A.6 Characterization and Results

A.6.1. Effect on surface finish

Surface finish and profile of the coated samples were checked using a Federal

Surface Analyzer system 4000 (Figure A.11) – which is capable of measuring the

roughness, waviness, form and profile of surfaces. Table below (Table A.4) shows the

surface finish of samples 2 -9 (Figure A.12-A.17). It was noted that the surface was

extremely rough (with surface finish ranging from 5 – 15 microns). Surface roughness,

however, smoothed out the gross waviness typical in these type of processes because of

166
cyclical deposition of weld beads. For this reason, the surface profile might be a better

indicator of the surface condition.

Figure A.11. Federal Surface Analyzer used in the surface measurement

Speed Power Hatch Spacing


# (m/sec) (Watts) (inches) Pt (µm) Ra (µm)
2 0.875 310 0.01 101.4 7.3
3 0.875 380 0.01 102.6 10.5
4 0.5 310 0.01 185.5 14.6
5 0.5 380 0.01 183.9 9.7
6 0.5 310 0.015 150 13.3
7 0.5 380 0.015 186.5 14.3
8 0.875 310 0.015 72.9 8.5
9 0.875 380 0.015 98.5 7.2

Table A.4. Measured surface roughness and profile of LENS coated NiAl samples

167
Figure A.12. Profile – sample 2 (in microns)

Figure A.13. Profile – sample 4 (in microns)

Figure A.14. Profile – sample 5 (in microns)

168
Figure A.15. Profile – sample 7 (in microns)

Figure A.16. Profile – sample 8 (in microns)

Figure A.17. Profile – sample 9 (in microns)

169
The data obtained from the surface roughness measurements was regressed to obtain the

following equation. The R-Square value obtained was .965 and the adjusted R-Square

value was .94, indicating a good fit. All three variables have an equal effect on the surface

profile.

Equation regressed:

Pt = 251.5-220.3 * TS (in m/sec) +.22* Power (watts) - 3275 * Hatch Spacing (in

thousandths).

A.6.2. Coating Morphology

To see effect of the process on the microstructure and the interface, all samples

were observed under a light microscope at 50X magnification. Figure A.19 shows the

results of the optical microscopy. Figure A.20 shows that the bond obtained is relatively

pore free, indicating excellent interface quality. Figure A.20 and Figure A.21 shows the

SEM performed on sample 4, indicating presence of several partially melted nickel

aluminide particles. Further analysis is required to relate the microstructure, the surface

quality and bond interface to the process variables.

170
Sample 2 Sample 3

Sample 4 Sample 5

Continued

Figure A.18. Optical images of coated surface at 50X for the samples 2-9. Note: samples

3, 5, 7, and 9 were coated at higher power levels, samples 6,7,8 and 9 have .015 HS

171
Figure A18 continued

Sample 6 Sample 7

Sample 8 Sample 9

172
a) b)

Figure A.19. SEM image showing cross section of sample 7, polished and bakelite

mounted. Sample was not etched a) 75X b)300X

Figure A.20. SEM image of surface of NiAl coating on H-13 substrate showing random

appearance of partly melted powder. Image was taken at 75X and 381X

173
a) b)

c) d)

Figure A.21. SEM Image of sample 7, showing NiAl powder at different stages of

melting. Images are at a)150X b)1000X c) 1200X and b) 5000X

174
Figure A.22. EDS analysis of a) Surface of sample 7 and b) NiAl powder

Table A.5 and Figure A.23 indicate the compositional gradient (sample 4) across

the H13 /NiAl interface. SEM indicates that the interface is sharp possibly because the

extremely high cooling rates. This may restrict the diffusion of Ni and Al into the steel

substrate due to lack of time at higher temperatures where diffusion is easier. The dilution

in the melt pool, correlated to the depth of melt pool and the table speed, is seen to be

uniform in the clad region.

175
6

Index Location from Fe (%) Ni (%) Al (%) Pb (%) Cr (%)


surface(in µm)
Surface 46.31 35.86 7.2
1 15 51.07 32.66 5.49 3.81 3.94
2 30 50.85 33.64 5.4 4.3 3.97
3 45 51.17 32.46 5.24 4.42 3.37
4 60 48.56 32.76 12.42 6.34 3.55
5 90 54.22 33.17 5.26 0 4.51
6 125 51.58 34.09 5.78 3.09 3.78
7 155 83.98 1.19 1.14 2.74 6.1
8 187 83.6 1.17 1 0 7

Table A.5. Compositional gradient (wt %) from surface to the bulk of substrate, obtained

through Energy Dispersive Spectrometric (EDS) analysis

176
90

80

70

60
Weight %

Fe (%)
50
Ni (%)
40 Cr (%)
30

20

10

0
0 15 30 45 60 90 125 155 187
Location (in microns)

Figure A.23. Variation of nickel, iron and chromium across the interface

177
REFERENCES

[1] Arai, T., "Tool materials and surface treatments," Journal of Materials Processing
Technology, Vol. 35, pp. 515-528, 1992.

[2] Arai, T., "Research and application of carbide and nitride coatings on to aluminum
die casting molds in Japan," presented at Die Casting Innovation, Indianapolis,
Indiana, 1995.

[3] Arai, T., Fujita, H., Sugimoto, Y., and Ohta, Y., "Diffusion carbide coatings
formed in molten borax systems," presented at Surface Modifications and
Coatings, Toronto, Ontario, Canada, 1995.

[4] Arai, T. and Iwama, T., "Carbide surface treatment of die cast dies and die
components," presented at 11th International Die Casting Congress and
Exposition, Cleveland, Ohio, 1981.

[5] Arai, T. and Komatsu, N., "Carbide coating process by use of salt bath and its
application to metal forming dies," pp. 225-231, 1982.

[6] Asm, Metals handbook: Heat treating, Cleaning, and Finishing, Vol. 2. Metals
Park, Ohio: ASM, 1964.

[7] Aston, J. and Barry, E., "A further consideration of factors affecting the life of
drop forging dies," Journal of the Iron and Steel Institute, No. July, pp. 520-6,
1972.

[8] Aston, J. L., Hopkins, A. D., and Kirkham, K. E., "The Wear testing of hot work
die steels," Metallurgia and Metal Forming, No. Feb., pp. 46-50, 1972.

[9] Banerjee, R., Collins, P. C., and Fraser, H. L., "Laser Deposition of In Situ Ti and
TiB Composites," Advanced Engineering Materials, Vol. 4, No. 11, pp. 847 -
851, 2002.

178
[10] Bariani, F. P., Berti, A. G., Luciano, D. A., and Roberto, G., "Wear in Hot and
Warm Forging: Design and Validation of a New Laboratory Test," Annals of the
CIRP, Vol. 45, pp. 249-253, 1996.

[11] Batit, G., Renaudin, J., Thore, Y., and Felder, E., "A Way to Build a Computer
Program for Forecasting the Abrasive Wear of Hot Forging Dies," presented at
11th International Drop Forging Congress, Cologne, Germany, 1983.

[12] Boddu, M. R., Landers, R. G., and Liou, F. W., "Control of Laser Cladding For
Rapid Prototyping – A Review," presented at Solid Freeform Fabrication
Proceedings, Austin, TX, 2001.

[13] Bramley, A., Lord, J., and Beeley, P., "Determination of Wear Resistance of Hot
Work Die Materials," Annals of the CIRP, Vol. 38/1, pp. 231-234, 1989.

[14] Burgreev, V. S. and Dobar, S. A., "Electolytic boriding of hammer forging dies
and their heat treatment," Metal Science and Heat treatment, Vol. 14, No. 6, pp.
513-517, 1972.

[15] Carpenter, "Series of Brochures from Carpenter Specialty Alloys."


www.cartech.com, 1999.

[16] Collins, P. C., Banerjee, R., and Fraser, H. L., "The influence of the enthalpy of
mixing during the laser deposition of complex titanium alloys using elemental
blends," Scripta Materialia, Vol. 48, No. 10, pp. 1445-1450, 2003.

[17] Cser, L., Geiger, M., Lange, K., Kals, J., and Hansel, M., "Tool Life and Tool
Quality in Bulk Metal Forming," Proceedings of Mechanical Engineers, Vol. 207,
pp. 223-239, 1993.

[18] Dean, T., "Wear in Drop Forging Dies," presented at 15th International MTDR
Conference, Birmingham, England, 1974.

[19] Dean, T. A., "A Comparisn of High Rate and Conventional Forging Machines,"
presented at 7th International MTDR Conference, University of Birmingham,
1966.

[20] Dean, T. A., "The net-shape forming of gears," Materials and Design, Vol. 21,
No. 4, pp. 271-278, 2000.

[21] Dean, T. A. and Sturgess, C., "Warm Forming Practice," Journal of Mechanical
Working Technology, Vol. 2, pp. 255-259, 1987.

179
[22] Deevi, S. C. and Sikka, V. K., "Nickel and iron aluminides: An overview on
properties, processing, and applications," Intermetallics, Vol. 4, No. 5, pp. 357-
375, 1996.

[23] Dennis, J. K. and Mahmoud, E. A. A. G., "Wear Resistance of Surface-treated hot


forging dies," Tribology International, Vol. 20, No. No 1 Feb., pp. 10-17, 1987.

[24] Dieter, G. E., Mechanical Metallurgy, 3rd ed. New York: McGraw-Hill, Inc.,
1986.

[25] Doege, E., "Aspects of Wear Prediction in Precision Forging," Proceedings of


Institiute of Mechanical Engineers, Vol. 208, pp. 111-119, 1994.

[26] Doege, E., Melching, R., and Kowallick, G., "Investigation into the Behavior of
Lubricants and the Wear Resistance of Die Materials in Hot and Warm Forging,"
J. Mech. Working Technology, Vol. 2, pp. 129, 1978.

[27] Doege, E. and Nagele, H., "FE Simulation of the Precision Forging Process of
Bevel Gears," Annals of the CIRP, Vol. 43, pp. 241-244, 1994.

[28] Doege, E., Romanowski, C., and Seidel, R., "Increasing tool life quantity in die
forging: chances and limits of tribological measures," presented at 24th NAMRC
Conference, Ann Arbor, MI, 1996.

[29] Douglas, R. and Kuhlmann, D., "Guidelines for precision hot forging with
applications," Journal of Materials Processing Technology, Vol. 98, No. 2, pp.
182-188, 2000.

[30] Edenhofer, B., "Production Ion nitriding," Metal Progress, Vol. 109, No. 03, pp.
38, 1976.

[31] Farmer, H. N., "Factors Affecting Selection and Performance of Hard-facing


Alloys," in Source Book on Wear Control Technology. Ontario, Canada: ASM,
1979.

[32] Felder and Montagut, "Friction and Wear During the Hot Forging of Steels,"
Tribology International, No. April, 1980, pp. 62-68, 1980.

[33] Felder, E. and Montagut, J., "Hot Forging of Steels," Tribology International, No.
April, pp. 61-68, 1980.

[34] Fiedler, H. C. and Dobnar, S. A., "Boriding steels for wear resistance," Metal
Progress, Vol. 99, No. 02, pp. 101, 1972.

[35] Fischer, P., Karapatis, N., Romano, V., Glardon, R., and Weber, H. P., "A model
for the interaction of near-infrared laser pulses with metal powders in selective
180
laser sintering," Applied Physics A Materials Science & Processing, Vol. 74, No.
4, pp. 467 - 474, 2002.

[36] Fischer, P., Romano, V., Weber, H. P., Karapatis, N. P., Boillat, E., and Glardon,
R., "Sintering of commercially pure titanium powder with a Nd:YAG laser
source," Acta Materialia, Vol. 51, No. 6, pp. 1651-1662, 2003.

[37] Fu, Y., Loredo, A., Martin, B., and Vannes, A. B., "A theoretical model for laser
and powder particles interaction during laser cladding," Journal of Materials
Processing Technology, Vol. 128, No. 1-3, pp. 106-112, 2002.

[38] Griffith, M. L., Ensz, M. T., Puskar, J. D., Robino, C. V., Brooks, J. A., Philliber,
J. A., Smugeresky, J. E., and Hofmeister, W. H., "Understanding the
microstructure and properties of components fabricated by Laser Engineered Net
Shaping (LENS)," Materials Research Society Symposium - Proceedings, Vol.
625, pp. 9-20, 2000.

[39] Grujicic, M., Hu, Y., Fadel, G. M., and Keicher, D. M., "Optimization of the
LENS Rapid Fabrication Process for In-Flight Melting of Feed Powder," Journal
of Materials Synthesis and Processing, Vol. 9, No. 5, pp. 223-233, 2001.

[40] Hu, Y. P., Chen, C. W., and Mukerjee, K., "Development of a new laser cladding
process for manufacturing cutting and stamping dies," Journal of Materials
Science, Vol. 33, No. 5, pp. 1287-1292, 1998.

[41] J. A. Brooks, C. V. Robino, T. Headley, S. Goods, R. C. Dykhuizen, and M. L.


Griffith, "Microstructure and Property Optimization of LENS® Deposited H13
Tool Steel," presented at Solid Freeform Fabrication Symposium, Austin, TX,
1999.

[42] Jeong, D. J., Kim, D. J., Kim, J. H., Kim, B. M., and Dean, T. A., "Effects of
surface treatments and lubricants for warm forging die life," Journal of Materials
Processing Technology, Vol. 113, No. 1-3, pp. 544-550, 2001.

[43] Kang, J. H., Park, I. W., Jae, J. S., and Kang, S. S., "A study on a die wear model
considering thermal softening: (I) Construction of the wear model," Journal of
Materials Processing Technology, Vol. 96, No. 1-3, pp. 53-58, 1999.

[44] Kang, J. H., Park, I. W., Jae, J. S., and Kang, S. S., "A study on die wear model
considering thermal softening (II): Application of the suggested wear model,"
Journal of Materials Processing Technology, Vol. 94, No. 2-3, pp. 183-188,
1999.

[45] Kar, A. and Mazumder, J., "Laser chemical vapor deposition of thin films,"
Materials Science & Engineering, Vol. B41, No. 3, pp. 368-373, 1996.

181
[46] Kathuria, Y. P., "Laser-cladding process: a study using stationary and scanning
CO2 laser beams," Surface and Coatings Technology, Vol. 97, No. 1-3, pp. 442-
447, 1997.

[47] Ki, H., Mohanty, P. S., and Mazumder, J., "Modelling of high-density laser-
material interaction using fast level set method," Journal of Physics D: Applied
Physics, Vol. 34, No. 3, pp. 364-372, 2001.

[48] Krauss, G., Steels: Heat Treatment and Processing Principles, 4th ed. Materials
Park, OH: ASM, 1995.

[49] Krishnadev, M. R. and Jain, S., "Enhancing Hot Forging Die life," Forging, No.
Winter, pp. 67-72, 1997.

[50] Lewis, G. K. and Schlienger, E., "Practical considerations and capabilities for laser
assisted direct metal deposition," Materials and Design, Vol. 21, No. 4, pp. 417-
423, 2000.

[51] Li, Y. and Ma, J., "Study on overlapping in the laser cladding process," Surface
and Coatings Technology, Vol. 90, No. 1-2, pp. 1-5, 1997.

[52] Lin, J. and Hwang, B.-C., "Clad profiles in edge welding using a coaxial powder
filler nozzle," Optics and Laser Technology, Vol. 33, No. 4, pp. 267-275, 2001.

[53] Liu, C.-C. and Huang, J.-L., "Effect of the electrical discharge machining on
strength and reliability of TiN/Si3N4 composites," Ceramics International, Vol.
29, No. 6, pp. 679-687, 2003.

[54] Mahrle, A. and Schmidt, J., "The influence of fluid flow phenomena on the laser
beam welding process," International Journal of Heat and Fluid Flow, Vol. 23,
No. 3, pp. 288-297, 2002.

[55] Manson, S. S., "Predictive Analysis of Metal Fatigue in The High Cycle Life
Range," presented at Fatigue at Elevated Temperatures, Storrs, Conn., 1972.

[56] Miller, R. S., Cao, G., and Grujicic, M., "Monte Carlo Simulation of Three-
Dimensional Nonisothermal Grain-Microstructure Evolution: Application to LENS
Rapid Fabrication," Journal of Materials Synthesis and Processing, Vol. 9, No. 6,
pp. 329-345, 2001.

[57] Ming, Q., Lim, L. C., and Chen, Z. D., "Laser cladding of nickel-based hardfacing
alloys," Surface and Coatings Technology, Vol. 106, No. 2-3, pp. 174-182, 1998.

[58] Netthofel, F., "Contributions to the Knowledge on Wear in Die Materials (in
German)." Hannover: Technical University of Hannover, 1965.

182
[59] Nugent, R. M., "Alloy 625 Surfacing of Tool and Die Steels," Welding Journal,
Vol. 5, No. No 6, June, pp. 33-39, 1986.

[60] Oddy, A. S., "Numerical prediction of microstructure and hardness in multicycle


simulation," Journal of Materials Engineering and Performance, Vol. 5, No. 3,
pp. 365

365-372, 1996.

[61] Okell, R. E. and Wolstencroft, F., "A Suggested Mechanism of Hot Forging Die
Failure," Metal Forming, Vol. Feb., pp. 41-9, 1968.

[62] Painter, B., Shivpuri, R., and Altan, T., "Prediction of Die Wear During Hot-
Forging of Automotive Exhaust Valves," Journal of Materials Processing
Technology, Vol. 59, pp. 132-143, 1996.

[63] Pickering, F. B., "The Properties of Tool Steels for Mould and Die Applications,"
presented at Tool Materials for Molds and Dies Application and Performance,
Golden, CO, 1987.

[64] Qian, M., Lim, L. C., Chen, Z. D., and Chen, W. L., "Parametric Studies of Laser
Cladding Processes," Journal of Materials Processing Technology, Vol. 63, No.
1-3, pp. 590-593, 1997.

[65] Roberts, G., Krauss, G., and Kennedy, R., Tool Steels, 5th ed. Metals Park, OH:
ASTM, 1998.

[66] Saiki, H., Marumo, Y., Minami, A., and Sonoi, T., "Effect of the surface structure
on the resistance to plastic deformation of a hot forging tool," Journal of
Materials Processing Technology, Vol. 113, No. 1-3, pp. 22-27, 2001.

[67] Schlienger, M. E., Harwell, L. D., Oliver, M. S., Baldwin, M. D., Ensz, M. T.,
Essien, M., Brooks, J., Robino, C. V., Smugeresky, J. E., Hofmeister, W. H.,
Wert, M. J., and Nelson, D. V., "Understanding thermal behavior in the LENS
process," Materials and Design, Vol. 20, No. 2-3, pp. 107-113, 1999.

[68] Sexton, L., Lavin, S., Byrne, G., and Kennedy, A., "Laser cladding of aerospace
materials," Journal of Materials Processing Technology, Vol. 122, No. 1, pp. 63-
68, 2002.

[69] Sharma, R. and Arrowsmith, D., "The Wear of Forging Dies in the First Five
Forging Blows," Wear, Vol. 74, pp. 1-10, 1981.

183
[70] Shi, J. and Liu, C. R., "Flow stress property of a hardened steel at elevated
temperatures with tempering effect," International Journal of Mechanical
Sciences, Vol. 46, No. 6, pp. 891-906, 2004.

[71] Shivpuri, R. and Semiatin, L., "Wear of Dies and Molds, and Coatings for Wear
Resistance," Department of Industrial Engineering ; Ohio State University,
Columbus, Ohio ERC/NSM-88-05, June 1988.

[72] Silva, T. M. and Dean, T. A., "Wear in drop forging dies," presented at 15th
International MTDR Conference, Birmingham, England, 1971.

[73] Singh, A. K., Rooks, B. W., and Tobias, S. A., "Factors affecting die wear," Wear,
Vol. 25, pp. 271-79, 1973.

[74] Stachowiak, G. W., Engineering tribology. New York: Elsevier, 1993.

[75] Subramanian, C., Strafford, K. N., Wilks, T. P., and Ward, L. P., "On the design
of coating systems: metallurgical and other considerations," Proceedings of the
1993 International Conference on Advances in Material & Processing
Technologies, AMPT'93, Vol. 56, No. 1-4, pp. 385-397 1996.

[76] Summerville, E., Venkatesan, K., and Subramanian, C., "Wear processes in hot
forging press tools," Materials and Design, Vol. 16, No. 5, pp. 289-294, 1995.

[77] Taylor, A. M. and Tookey, K. M., "Ion-Nitriding - a Process With a Future,"


Metallurgia, Vol. 48, No. 02 Feb., pp. 64, 1981.

[78] Thomas, A., "Wear of Drop Forging Dies," Tribology in Iron and Steel Works
(Iron and Steel Institute), 1970.

[79] Tittagala, S. R., Beeley, P. R., and Bramley, A. N., "Wear Test for Hot-working
Die Steels," Metals Technology, Vol. 9, No. Nov., pp. 434-439, 1982.

[80] Tsuchiya, Y., Kawaura, H., Hashimoto, K., Inagaki, H., and Arai, T., "Core Pin
Failure in Aluminum Die Casting and Surface Treatment," presented at The Many
Faces of the Die Casting and the Effect of Surface Treatment, Minneapolis,
Minnesota, 1997.

[81] Tulsyan, R., Shivpuri, R., and Altan, T., "Investigation of Die Wear in Extrusion
and Forging of Exhaust Valves," ERC/NSM - Ohio State University, Columbus,
Ohio B-93-28, August, 1993 1993.

[82] Vasinonta, A., Beuth, J. L., and Griffith, M. L., "Process Maps for Controlling
Residual Stress and Melt Pool Size in Laser-based SFF Processes," presented at
Solid Freeform Fabrication Proceedings, Austin, TX, 2000.

184
[83] Venkatesan, K., Summerville, E., and Subramanian, C., "Performance of Surface
Engineered Hot Forging Dies," Materials Australia, Vol. 30, No. 3, pp. 10-12,
1998.

[84] Weist, C. and Westheide, C. W., "Application of Chemical and Physical Methods
for the Reduction of Tool Wear in Bulk Metal Forming Processes," Annals CIRP,
Vol. 35, No. 01, 1986.

[85] Yevko, V., Park, C. B., Zak, G., Coyle, T. W., and Benhabib, B., "Cladding
formation in laser-beam fusion of metal powder," Rapid Prototyping Journal, Vol.
4, No. 4, pp. 168-184, 1998.

185

You might also like