Вы находитесь на странице: 1из 281

About the Book

Conceived from years of problem solving,


Tool and Die Making Troubleshooter is
destined to be an indispensable guide for

Tool and Die Making Troubleshooter


designing, constructing, and maintaining
tools, dies, molds, and fixtures. The book
contains hands-on information, valuable
tooling tips, and procedural recommendations
regarding the selection, processing, and use of materials. This com-
prehensive reference is for anyone working with the selection, ap-
plication, and use of cast irons, cast steels, wrought tooling sheets,
and aluminum and stainless steels. The subjects of basic metallurgy,
heat treating, machining, grinding, electrical discharge machining,
Tool and
welding, quality, wear enhancements, surface coatings, and treat-
ments for tools and dies have been carefully organized and pre-
sented for quick reference.
Die Making
About the Author
Richard Leed began his career with Leed
Steel Company, a family-owned metal
service center. In 1980, he became Vice
TROUBLESHOOTER
President and quickly realized the im-
portance of cost-effective tooling and its
impact on part quality and the financial
success of a production operation. With this in mind, he took a
strong interest in providing his customers with the technical and
metallurgical assistance needed to optimize their tooling invest-
ments. These services included technical discussions, training pre-
sentations, and specification and procedure upgrades for tool design,
material selection, machining, heat treating, finish machining
(grinding, welding, and electrical discharge machining), tool and
die failure analysis, and troubleshooting. Richard Leed is a tool
and die troubleshooting consultant to industry. He has authored
numerous technical papers and manuals that focus on improving
tool and die performance.
Richard M. Leed

Leed
Society of Society of
Manufacturing Manufacturing
Engineers Engineers
www.sme.org www.sme.org

Association for Association for


Forming & Fabricating Forming & Fabricating
Technologies/SME Technologies/SME
www.sme.org/afft www.sme.org/afft
Tool and Die Making
Troubleshooter

prepages.pmd 1 10/29/02, 11:10 AM


prepages.pmd 2 10/29/02, 11:10 AM
Tool and Die Making
Troubleshooter

Richard M. Leed

Society of Association for


Manufacturing Forming & Fabricating
Engineers Technologies/SME
Dearborn, Michigan

prepages.pmd 3 10/29/02, 11:10 AM


Copyright © 2003 by the Society of Manufacturing Engineers

987654321

All rights reserved, including those of translation. This book, or parts


thereof, may not be reproduced by any means, including photocopying,
recording or microfilming, or by any information storage and retrieval
system, without permission in writing of the copyright owners.

No liability is assumed by the publisher with respect to use of informa-


tion contained herein. While every precaution has been taken in the
preparation of this book, the publisher assumes no responsibility for
errors or omissions. Publication of any data in this book does not consti-
tute a recommendation or endorsement of any patent, proprietary right,
or product that may be involved.

Library of Congress Catalog Card Number: 2002113502


International Standard Book Number: 0-87263-643-7

Additional copies may be obtained by contacting:


Society of Manufacturing Engineers
Customer Service
One SME Drive, P.O. Box 930
Dearborn, Michigan 48121
1-800-733-4763
www.sme.org

SME staff who participated in producing this book:


Phil Mitchell, Senior Editor
Eugene Sprow, Editor
Rosemary Csizmadia, Production Supervisor
Frances Kania, Production Assistant
Kathye Quirk, Graphic Designer/Cover Design
Jon Newberg, Production Editor

Printed in the United States of America

prepages.pmd 4 10/29/02, 11:10 AM


About The Society of Manufacturing Engineers (SME)
The Society of Manufacturing Engineers is the world’s leading
professional society supporting manufacturing education. Through
its member programs, publications, expositions, and professional
development resources, SME promotes an increased awareness
of manufacturing engineering and helps keep manufacturing pro-
fessionals up to date on leading trends and technologies. Head-
quartered in Michigan, SME influences more than half a million
manufacturing engineers and executives annually. The Society has
members in 70 countries and is supported by a network of hun-
dreds of chapters worldwide. Visit us at www.sme.org.

About AFFT/SME
The Association for Forming & Fabricating Technologies of SME
(AFFT/SME) focuses on the technologies and processes that effi-
ciently make products from metal sheet, coil, plate, tube, or pipe
stock. Typical industries served are automotive, off-highway, aero-
space, defense, appliance, furniture, and consumer electronics
products. Core processes include general pressworking—stamp-
ing, drawing, forming, bending, and shearing, for example—as
well as the fabricating technologies of punching, cutting, sawing,
welding, and others.
Many AFFT/SME members are manufacturing, tool, or pro-
cess engineers specializing in keeping production forming or fab-
ricating technologies current for their companies; general
stamping, fabricating, welding, and assembly managers respon-
sible for overseeing plant operations; and owners, partners, pro-
prietors, and other company officials of relevant job shops. AFFT/
SME membership allows such people to more quickly identify in-
novations that can lower costs while increasing product quality
and yield. The AFFT/SME community also fosters learning among
members to better manage their businesses and interact with
customers, suppliers, and industry partners.

prepages.pmd 5 10/29/02, 11:10 AM


prepages.pmd 6 10/29/02, 11:10 AM
This is dedicated to my wife, Nancy, and my children
Richard and Carolyn. Together they gave me purpose,
taught me the meaning of unqualified love,
and gave me inspiration for the future.

prepages.pmd 7 10/29/02, 11:10 AM


Tool and Die Making Troubleshooter

viii

prepages.pmd 8 10/29/02, 11:10 AM


Chapter 1: The Tooling Investment

Table of Contents

Preface ..................................................................................... xiii


Acknowledgments .................................................................... xv
1 The Tooling Investment ........................................................ 1
Tooling Facts of Life ........................................................................ 1
Time ................................................................................................ 3
2 Steelmaking Methods and Hot Working .............................. 7
The Origins of Steelmaking ............................................................ 7
The Legend of Damascus Steel ..................................................... 9
Electric Furnace Melting ............................................................... 10
Quality Effects of Ingot Mold Solidification .................................. 12
Refining/Remelting Processes ..................................................... 14
Powder-metal Tool Steels ............................................................. 19
Hot Working of Tool Steels ........................................................... 21
Cast-to-shape Versus Wrought Tool Steels ................................. 24
3 Quality Considerations ....................................................... 27
ISO 9000 Quality Systems ........................................................... 27
Aircraft-quality Steel ..................................................................... 30
Mold-quality Steel ......................................................................... 31
4 Metallurgy and Engineering Considerations .................... 35
Basic Mechanical Properties ........................................................ 35
Hydrogen Embrittlement .............................................................. 39
Alloy Segregation (Banding) ........................................................ 40

ix

prepages.pmd 9 10/29/02, 11:10 AM


Tool and Die Making Troubleshooter

The Art of Spark Testing ............................................................... 42


Effects of Alloying Elements ......................................................... 43
Crystal Structures in Heat Treatment ........................................... 45
Effects of Carbides ....................................................................... 47
Directional Properties ................................................................... 47
Wear and Fatigue ......................................................................... 48
Tool-steel Toughness .................................................................... 51
Thermal Conductivity and Fatigue ............................................... 53
Surface Scale and Decarburization ............................................. 57
Grain Coarsening from Overheating ............................................ 60
Incipient Melting ........................................................................... 61
Retained Austenite ....................................................................... 63
Stress-related Problems ............................................................... 64
Orange Peel and Pitting ............................................................... 65
5 Tooling Material Selection .................................................. 69
Identifying Tool-steel Grades ........................................................ 69
Tool-steel Classifications .............................................................. 70
Tool-steel Selection ...................................................................... 75
6 Tooling Design .................................................................... 91
More Than just Making Prints ...................................................... 91
Common Design Faults ................................................................ 92
Common Failure Modes ............................................................... 94
7 Tool Machining and Welding ............................................ 109
Machinability ............................................................................... 109
Tool-wear Classifications ............................................................ 120
EDM Effects on Tool Steels ........................................................ 130
Welding Tool Steels .................................................................... 136
Shrink Fitting ............................................................................... 142
8 Heat Treatment .................................................................. 147
More than just Hardness ............................................................ 147
Leave Nothing to Chance ........................................................... 150
Furnace Loading ........................................................................ 152
Hardening and Related Stresses ............................................... 156
Stainless Foil Wraps ................................................................... 158
Gas Atmospheres ....................................................................... 159
Vacuum Heat Treatment ............................................................. 161

prepages.pmd 10 10/29/02, 11:10 AM


Table ofInvestment
Chapter 1: The Tooling Contents

Annealing Alloy and Tool Steels ................................................. 163


Normalizing ................................................................................. 164
Stress Relieving .......................................................................... 166
Preheat Before Hardening ......................................................... 168
Hardness-caused Cracking ....................................................... 169
Austenitizing Soaking Times ...................................................... 172
Quenching .................................................................................. 173
Tempering ................................................................................... 179
Subzero Treatments ................................................................... 180
Distortion and Size Change ....................................................... 183
Troubleshooting .......................................................................... 186
9 Hardness Testing .............................................................. 193
File Testing .................................................................................. 193
Rockwell Testing ......................................................................... 195
Brinell Testing ............................................................................. 196
10 Wear Enhancement Treatments ....................................... 203
Surface Treatments are not Coverups ....................................... 203
A Firm Foundation ...................................................................... 204
Progress in Wear Enhancement ................................................ 205
Appendix A: Material Data .................................................... 217
A2 Tool Steel ............................................................................... 217
D2 Tool Steel ............................................................................... 219
H13 Tool Steel ............................................................................. 220
M2 High-speed Steel .................................................................. 222
M4 High-speed Steel .................................................................. 224
O1 Tool Steel .............................................................................. 226
S5 Tool Steel ............................................................................... 227
S7 Tool Steel ............................................................................... 229
W1 Carbon Tool Steel ................................................................. 230
420 Martensitic Stainless Steel .................................................. 232

Appendix B: Glossary ............................................................. 235


Index ........................................................................................ 257

xi

prepages.pmd 11 10/29/02, 11:10 AM


Tool and Die Making Troubleshooter

xii

prepages.pmd 12 10/29/02, 11:10 AM


Chapter 1: The Tooling Investment

Preface

Quality dies, molds, fixtures, machine parts, and cutting tools


are the very heart of manufacturing operations. Everyone who
constructs, uses, and/or maintains these components must ensure
that their quality is sufficient and cost effective for meeting pro-
duction requirements. Construction problems, misuse, and rushed
maintenance procedures are often the cause of very expensive pro-
duction delays and poor part quality.
The Tool and Die Making Troubleshooter is a comprehensive
reference book for every individual working with the selection,
application, and use of steels for tools, dies, and molds. This book
contains information, comparison charts, technical data, tooling
tips, and procedural recommendations about material selection,
processing, and use. It progressively describes the metallurgy, heat
treating, machining, grinding, electrical discharge machining,
welding, wear-enhancement coatings, and treatments for tool and
die steels.
It is the intention that this book become an indispensable guide
for designing, constructing, and maintaining quality tools, dies,
molds, and fixtures.
You may contact the author at the following address: Richard
M. Leed, Leed Steel Co., Inc., 228 Sawyer Ave., Tonawanda, NY
14150; phone: 716-874-2554; fax: 716-874-2438; E-mail:
rleed@leed steel.com.

xiii

prepages.pmd 13 10/29/02, 11:10 AM


Tool and Die Making Troubleshooter

xiv

prepages.pmd 14 10/29/02, 11:10 AM


Chapter 1: The Tooling Investment

Acknowledgments

Many people helped me write this book. Their expertise, guid-


ance, and support were critical in making it accurate and compre-
hensive. The following people deserve special recognition:
Mr. William D. Leed, Jr. was founder of Leed Steel Company.
Dad gave me the opportunity to work in this wonderful field. He
taught me how to laugh at myself as well as a good joke.
My mother Myra kept my family bonded together. She taught
me to value hard work and family ties.
Mr. Thomas Leed, my brother and partner, shares the same
goal of an honest and healthy partnership.
Mr. Theodore Megas was a Metallurgist and Sales Engineer with
Bethlehem Steel Company. Before his death, Ted was my mentor.
He taught me how to patiently deal with difficult situations and
earn a customer’s respect.
Mr. Browney Czajka was a Tool and Die Engineer at Ford Mo-
tor. Browney was a man who let his experience, deeds, and ac-
tions speak louder than his words.
Mr. Frank Donchez, who is a tool steel expert, tool and die con-
sultant, and trustworthy friend, and is now retired from Bethlehem
Steel Company. Frank taught me how to share knowledge and
deal with others patiently, accurately, and with integrity.

xv

prepages.pmd 15 10/29/02, 11:10 AM


Tool and Die Making Troubleshooter

xvi

prepages.pmd 16 10/29/02, 11:10 AM


Chapter 1: The Tooling Investment

1
The Tooling Investment

The tooling investment begins with design and material


selection and moves on to encompass machining, heat treating,
finish machining, tryout, production, maintenance, and repair.
The success of the tooling investment requires balancing the
total cost in labor and materials of producing and maintaining
the tool against the return that tool achieves in production. A
quality tool produces quality parts and does so reliably over its
total useful life. Clearly, the most expensive tool is the tool that
fails prematurely.
This chapter examines the importance of the tooling invest-
ment. It outlines some of the more common causes of premature
failures. It also emphasizes how expensive these failures can be
in terms of the high costs of part repair, maintenance, production
downtime, and part quality.

TOOLING FACTS OF LIFE


Cost
The cost of tooling materials—specialty steels and alloys—is
insignificant when compared to the total costs of manufacturing
finished tools, dies, molds, machine parts, or fixtures. The high
cost of production downtime due to premature tool failure will
quickly negate any possible savings from choosing a less expensive
material.

Ch01.pmd 1 10/29/02, 11:10 AM


Tool and Die Making Troubleshooter

Teamwork
Designing, building, and maintaining production-worthy tools
and dies requires engineers, machinists, toolmakers, heat treaters,
and production personnel working together as a team.

Experience
Tool design involves more than just creating a tool and making
a print. Today’s tool-design engineer must be proactive—an
experienced tool-and-die troubleshooter who knows all the facets
of the toolmaking process and tool use.

Machining
Machining tooling to finish dimensions before hardening is a
dangerous practice. Size change and/or distortion can occur during
hardening, making it impossible to true up the part to the required
dimensions during final machining.

Heat Treatment
Every year valuable tools, dies, and molds—carefully designed
and machined—are sent to the heat treater with inadequate spec-
ifications. Heat treaters should be given the specific instruction
and comments necessary to accurately define tooling goals and
requirements.
Too much emphasis is placed on heat treating for hardness alone.
The hardening operation should be used to develop and fine tune
the working hardness levels with the different engineering and
physical properties needed to guarantee optimum tool performance.

Tempering
Sufficient tempering of as-quenched tooling is necessary to
achieve specific hardness levels, for stress relief, to develop desired
physical properties, and to promote dimensional stability. Never
short cycle the tempering operation.

Ch01.pmd 2 10/29/02, 11:10 AM


Chapter 1: The Tooling Investment

Proper ty Changes
Tooling personnel must be aware that heat generated during
grinding, welding, and electrical discharge machining (EDM)
frequently causes metallurgical and physical property changes that
adversely affect the service life of tooling.

Labor
Labor hours saved by short-cycling or speeding up established
procedures are too often lost when tools and dies fail prematurely
due to less than optimum practices in their construction or man-
ufacture.

Handling
Tools properly handled in service last longer and produce better
parts. Clearances between mating tools and dies, press maintenance
and alignment, material feeds, magnetism, lubrication, and
residual stresses are all concerns that affect the service life of
tools and dies.

TIME
How many times has this manufacturing lament been repeated?
“There is never enough time to do it right, but always enough
time to do it over!” The M4-steel milling cutter in Figure 1-1
exemplifies this time-worn proverb. Six of these parts cracked
before they could be put into service. To meet production schedules,
all of the parts were subjected to short cycling and a speedup in
the heat-treatment process. After hardening, the tempering
operation was also short cycled to save more time. The labor hours
supposedly saved by short cycling and rushing heat treatment were
only a fraction of the hours required to remake these parts after
they cracked. The lesson is clear: there is never a substitute for
proper care and proactive planning.

Ch01.pmd 3 10/29/02, 11:10 AM


Tool and Die Making Troubleshooter

Figure 1-1. Two views of one of six M4 high-speed-steel milling cutters that
cracked prematurely due to improper heat treatment and tempering.

Ch01.pmd 4 10/29/02, 11:11 AM


Chapter 1: The Tooling Investment

The High Cost of Downtime


Quality tools and dies are the very heart of manufacturing. Tool-
and-die engineers, tool-and-die makers, heat treaters, and
supervisory personnel must all work together as a team to
maximize the quality and performance of the tooling they
manufacture, use, or maintain. This is the only way to minimize
the high cost of production downtime, poor part quality, and the
need to replace tooling.
The aluminum casting die in Figure 1-2 cost approximately
$18,500 to construct. Typical maintenance reconditioning (stress
relieving, machining, welding, rehardening, EDM, etc.) after an
initial production run can cost approximately $8,000 to perform.
A careful review and upgrade of this die’s manufacturing and
maintenance specifications and procedures by tool-and-die
engineers, tool-and-die makers, heat treaters, and supervisory

Figure 1-2. Teamwork on this aluminum casting die produced a threefold


increase in life of the part.

Ch01.pmd 5 10/29/02, 11:11 AM


Tool and Die Making Troubleshooter

personnel resulted in a threefold increase in the part’s production


performance. Thus, tens of thousands of dollars were saved by
minimizing the high costs of production downtime and die
maintenance.

Tooling Cost Breakdown


The graph in Figure 1-3 illustrates the approximate costs of
manufacturing an automotive stamping die. It shows how
insignificant the cost of the tool steel is compared to the overall
combined costs of manufacturing finished tools, dies, and molds.
Note that tool steel at 1% is far less than the other material,
process, and labor costs.

Figure 1-3. Relative tool and die costs for a large automotive stamping die.

Ch01.pmd 6 10/29/02, 11:11 AM


Chapter 2: Steelmaking Methods and Hot Working

2
Steelmaking Methods and Hot Working

Although iron and steelmaking began over 3,000 years ago, the
most significant advances in melting practices occurred in the 20th
century. In the early 1900s, tool steels were melted in electric-arc
furnaces under a double-slag process that resulted in very clean
steel.
Today, tool-steel melting technology has changed, with the elec-
tric furnace used to melt the charge prior to secondary refining
by either argon-oxygen decarburization (AOD), vacuum-oxygen
decarburization (VOD), or other ladle refining processes. These
new practices provide assurance that the tool-steel product will
be very clean and more homogeneous when compared to other
more general steel products. Some tool steels, depending on their
end use, are further refined via consumable-electrode processes
to produce the ultimate in quality. This chapter addresses these
processes and answers some of the most common questions about
the manufacture of tool steels.

THE ORIGINS OF STEELMAKING


Archaeologists have found that as early as 4000 B.C. people were
using iron from meteorites to make weapons, ornaments, and tools.
No one knows for sure when or where man first smelted iron from
iron ore.

Ch02.pmd 7 10/29/02, 11:11 AM


Tool and Die Making Troubleshooter

Figure 2-1. Egyptian iron-smelting furnace (Stoughton 1934).

The Egyptian iron-smelting furnace shown in Figure 2-1 is from


a wall painting in the tomb of Rekhmara at Thebes, 18th Dynasty,
1535–1450 B.C.
The oldest known steel artifact was a tool found in a crevice in
one of the ancient Egyptian pyramids—a high-carbon steel dat-
ing back approximately 3,000 years. The oldest written reference
to steel is found in an ancient Chinese manuscript dating from
2500–1500 B.C. In this writing, a warlord spoke of a kind of com-
pass called a “south pointing chariot” used to aid his army in find-
ing their way home from a campaign far to the north. Scholars
believe that the Chinese must have used heat-treated steel for
their early compasses. Only heat-treated steel would be able to
develop a permanent magnetism. Iron would not have the needed
magnetic properties to act as a compass and point due north con-
sistently and accurately.
The Iron Age began in approximately 1500–1000 B.C. The ear-
liest iron to be worked by man was meteorites found in various
locations in India, Greece, Persia, and Egypt. Heated and pounded
into tools and weapons, this iron was known as the “gift from the

Ch02.pmd 8 10/29/02, 11:11 AM


Chapter 2: Steelmaking Methods and Hot Working

gods” because it mysteriously fell from the sky. Iron that was later
mined was known as lodestone. Only kings and powerful lords
could afford tools and weapons made from lodestone.

THE LEGEND OF DAMASCUS STEEL


Some of the earliest references to steel have been traced to the
Middle East and India. A steel from India called “wootz” was highly
prized and actively traded in and around the city of Damascus in
ancient Syria, where it became known as “Damascus steel.” It
was probably the early Greeks, 400 B.C., who began to earnestly
manufacture quantities of high-quality steel swords. It might have
been a fine Damascus steel sword that helped Alexander the Great
defeat the great Persian and Egyptian armies and conquer the
known world in 325 B.C. More frequent references to iron mak-
ing and steelmaking are found dating from 1000 B.C.–500 A.D.
The Romans made good use of iron and steel; Julius Caesar was
known to have had some chariots equipped with iron wheels.
The legend of Damascus steel was born in the Middle East in
the 11th through 13th centuries A.D. during the Crusades, as the
Christians and Muslims fought each other for more than 200 years.
The Christian Crusaders marveled at the ability of the swords
wielded by their Muslim adversaries to retain a keen cutting edge
and resist chipping, cracking, denting, or dulling from repeated
blows in combat. The swords of the Christians were cumbersome,
brittle, and unable to retain a sharp cutting edge when compared
to the Muslims’ dull-colored, strangely patterned steel blades.
In 1825, Sir Walter Scott popularized the legend of Damascus
steel in his novel, The Talisman. One story describes a meeting
between King Richard the Lionhearted and Saladin, the most noble
of his Muslim adversaries. While boasting, King Richard smashed
a steel mace in two with a tremendous blow from his heavy two-
handed broadsword.
Saladin’s response was to toss a silk scarf in the air, swiftly
draw his Damascus scimitar (curved sword) and slice it in half
without disturbing its descent to the ground. To quote Sir Walter
Scott, Saladin’s scimitar had “a curved and narrow blade which
glittered not like the swords of the Franks, but was on the con-

Ch02.pmd 9 10/29/02, 11:11 AM


Tool and Die Making Troubleshooter

trary, of a dull blue color, and marked with ten millions of mean-
dering lines.”
Whether or not the encounter between King Richard and
Saladin ever took place, the legend of Damascus steel grew. Ar-
morers, from as far back in time as Alexander were always very
secretive regarding their method of making Damascus steel. Mod-
ern-day metallurgists have unlocked those secrets. The key to
developing Damascus properties was in making a steel rich in car-
bon (1–2%), which was repeatedly forged and hammered at tem-
peratures in the range of 1,700° F (927° C). After forging to size,
the rough-shaped blades were reheated and rapidly cooled by
quenching. Legend has it that the best Damascus blades devel-
oped their strength from being quenched in the blood of a dragon
or in the bowels of a strong slave or brave captive.
Samurai swords of Japan are examples of the finest Damascus
steel ever produced. The extreme quality of Samurai Damascus
blades resulted from hammering out a bar to double its original
length, folding it over, then repeating the process as many as 1,000
times. In this way, two layers grew to four, four to eight, eight to
sixteen, and so on, until several thousand layers were hammered
into one single blade.
Following the Crusades, weapons of all kinds were fashioned
from Damascus steel. Today, metallurgists are optimistic that they
can apply some of the ideas of ancient Damascus manufacturing
and forging to modern-day metallurgy to make steels that are
tougher and stronger, with a better combination of impact strength
and wear resistance (Shelby and Wadsworth 1985; Sunday Demo-
crat and Chronicle 1981).

ELECTRIC FURNACE MELTING


In the past, tool steels were primarily produced in electric-arc
furnaces, which were usually round with carbon or graphite elec-
trodes extending through the roof (Figure 2-2).
The charge was a carefully measured and controlled combina-
tion of scrap and nonoxidizable ferro alloys. Electrodes lowered
into the furnace to a point near the charge melted the steel, and

10

Ch02.pmd 10 10/29/02, 11:11 AM


Chapter 2: Steelmaking Methods and Hot Working

Figure 2-2. Cross section of an electric furnace used for making tool steels
(United States Steel Corp. 1971).

an oxidizing slag was added to the molten heat to remove un-


wanted elements like phosphorus and sulfur. The oxidizing slag
was discarded and reducing slag added. The second slag process—
the reducing slag—served to enhance cleanliness and minimize
the loss of certain alloying elements that might burn off. Under
this slag, the balance of ferro alloys was added to adjust the heat
chemistry.
After the analysis of the molten steel had been checked, the
furnace was tapped. During tapping, the liquid steel was poured
into a ladle located in a pit at the back of the furnace. The molten
steel in the ladle was then poured into ingot molds and left to
solidify. This operation was called “teeming.” Once solidification
was complete, the steel ingots were stripped from their molds,
reheated, and hot-rolled into blooms or billets that were allowed
to cool. They were then carefully inspected and any surface im-
perfections were removed by scarfing and/or grinding. The blooms
or billets were heated once again and either forged or rolled into
bars or plates, and subsequently inspected and annealed.
Today, tool-steel melting technology has changed. Now the elec-
tric furnace is used only to melt the charge for secondary refining
via the argon-oxygen decarburization (AOD) or vacuum-oxygen

11

Ch02.pmd 11 10/29/02, 11:11 AM


Tool and Die Making Troubleshooter

decarburization (VOD) processes or other ladle refining tech-


niques. These secondary refining processes were developed to in-
crease the yield of the heat; to improve processing temperature
control, deoxidation, and microcleanliness; and to minimize im-
purities such as sulfur. The result is very clean, low-sulfur, homo-
geneous tool steels. Some tool steels, depending on their end use,
are further refined via consumable electrode processes such as
electroslag remelting (ESR) and vacuum arc remelting (VAR) to
produce the ultimate in quality (American Society for Metals 1948).

QUALITY EFFECTS OF INGOT MOLD SOLIDIFICATION


The solidification of molten tooling steels in ingot molds is a
critical factor governing their overall quality. Regardless of how it
is processed, no steel is completely free of nonmetallic inclusions
that become trapped as the molten steel solidifies in the ingot
mold. In the ingots shown in Figure 2-3, note that solidification

Figure 2-3. Ingot solidification stages during conventional steelmaking.

12

Ch02.pmd 12 10/29/02, 11:12 AM


Chapter 2: Steelmaking Methods and Hot Working

begins at the mold walls and gradually continues toward the cen-
ter. Nonmetallic inclusions in the steel tend to stay ahead of the
solidified layers and migrate toward the center of the ingot, then
up into the top of the ingot body.
Tool steels made by conventional (non-ESR and VAR) processes
are melted and then teemed into ingot molds to solidify. So-
lidification begins at the ingot mold walls and gradually contin-
ues toward the center. As the solidification process continues,
nonmetallic inclusions (impurities such as oxides, sulfides, and
silicates) in the steel tend to stay ahead of the solidified layers
and migrate toward the center of the ingot, then up into the top
of the ingot body. After solidification, portions of the ingot with
heavy concentrations of nonmetallic inclusions are discarded.
However, some nonmetallic inclusions may be trapped in the bal-
ance of the steel. When such inclusions are discovered, they will
most likely be found at a location corresponding to the metallur-
gical center of the ingot.
Figures 2-4 and 2-5 show a D2 tool-steel shaft with nonmetallic
inclusions discovered during the machining operation. Because of
this, a plastic mold designer would be well advised to keep “cen-
ter quality” in mind when designing a critical mold. For example,
when planning a 1-in. (25.4-mm) deep plastic mold cavity, a 2-in.

Figure 2-4. D2 tool-steel mixing shaft showed gross examples of nonme-


tallic inclusions during machining. Inclusions of this size and nature should
have been discovered by the manufacturer during ultrasonic testing.

13

Ch02.pmd 13 10/29/02, 11:12 AM


Tool and Die Making Troubleshooter

Figure 2-5. Closeup of the mixing shaft in Figure 2-4.

(50.8-mm) thick piece of stock would not be the best choice be-
cause this puts the bottom of the polished cavity at a depth statis-
tically most likely to have inclusions present. A better choice would
be thicker material (Allen 1969).

REFINING/REMELTING PROCESSES
The use of special refining and/or remelting processes for tool
steel, high-speed steels, stainless steel, and other high-alloy steels
is becoming increasingly popular because these techniques help
meet the growing demand for steels with improved mechanical
properties and cleaner, more uniform, and sounder internal struc-
tures. Argon-oxygen decarburization (AOD) and/or vacuum de-
gassing are used separately or at times in combination following
conventional electric-arc furnace melting. These processes may
be used to provide a finished product or in preparation for addi-

14

Ch02.pmd 14 10/29/02, 11:12 AM


Chapter 2: Steelmaking Methods and Hot Working

tional refining operations such as ESR, VAR, and vacuum induc-


tion melting (VIM) (United States Steel Corp. 1985).
Remelting is carried out in a “duplex” operation—two-furnace
melting. The first melting is usually performed in the electric-arc
furnace, and the remelting processes most often used are ESR,
VAR, and VIM.

Argon-oxygen Decarburization
In the AOD process, liquid metal obtained from conventional
electric-arc furnace melting or, in the case of small facilities, from
an induction furnace, is refined in a refractory-lined vessel by in-
jection of varying amounts of argon-oxygen gas mixtures. This pro-
cess was originally developed for the production of stainless steels,
but has been modified for the production of many other steels, in-
cluding tool steels. To accomplish the refining, the argon-oxygen
gas mixture is blown into the steel through an opening called a
“tuyère” located in the bottom of the AOD vessel. The blow is
usually done in several stages. The argon gas dilutes the carbon-
oxygen atmosphere in the melt. In the case of high-chromium
steels, this increases the affinity of carbon for oxygen, thus mini-
mizing the oxidation of chromium.
The AOD process has the economic benefit of reducing operat-
ing times at lower temperatures than would be necessary in the
electric-arc furnace. After processing in the AOD vessel, the steel
may be vacuum degassed and then conventionally teemed into
ingot molds. Experience has shown that the AOD process pro-
duces very clean tool steels—steel with a minimum of nonmetal-
lic inclusions, along with low oxygen content, which results in
cleaner steel. The process can also develop very low sulfur con-
tents when required.

Vacuum Degassing
Many tool steels and other high-alloy steels are subjected to a
vacuum degassing operation following their manufacture or dur-
ing processing in either a double-slag electric furnace or AOD ves-
sel. Vacuum degassing exposes molten steel to a low-pressure

15

Ch02.pmd 15 10/29/02, 11:12 AM


Tool and Die Making Troubleshooter

environment that reduces the gas in steel (for example, hydro-


gen, oxygen, and nitrogen).
The removal of hydrogen eliminates difficulties with its
embrittling effects. The removal of oxygen assists in reducing the
amount of oxide-type nonmetallic inclusions. Removal of both of
these gases plus nitrogen minimizes gas porosity. The resulting
steel is cleaner and more sound.

Electroslag Remelting
ESR is a secondary melting operation that has become one of
the most efficient ways to refine steels, making them cleaner and
more homogeneous and uniform in structure. The resulting high-
alloy tool steels are cleaner, tougher, and stronger. Figure 2-6 il-
lustrates this process.

Figure 2-6. The electroslag remelting furnace (Bethlehem Steel Corp. 1974).

16

Ch02.pmd 16 10/29/02, 11:12 AM


Chapter 2: Steelmaking Methods and Hot Working

In the ESR process, ingots are manufactured from steel melted


in electric-arc furnaces and vacuum degassed. The ingots are then
forged and/or turned into electrodes, which are carefully lowered
into a molten bath of slag inside the ESR furnace. Electric cur-
rent is passed through the steel electrode, the slag, and the re-
melted steel. As a result, the tip of the electrode melts and forms
tiny droplets of pure steel. As the tiny droplets travel through the
molten slag, a high percentage of impurities, such as sulfur and
other nonmetallic inclusions, are removed by chemical reaction.
The purified metal falls through a water-cooled section at the bot-
tom of the furnace and solidifies to form a new ingot.
The structure of the ESR-cast ingot is more uniform and ho-
mogeneous when compared to ingots produced from conventional
electric-arc melting and ladle casting. This is because the molten
bath of steel, which builds up drop by drop, solidifies more uni-
formly from bottom to top and from side to side. Centerline alloy
segregation and pipe are also dramatically reduced as compared
to conventional melting and casting techniques. In addition, the
steel is very low in nonmetallic inclusions.

Vacuum Arc Remelting


VAR is sometimes referred to as the consumable-electrode melt-
ing process. In the VAR furnace, steels produced in the electric-
arc furnace are remelted and poured into round ingots for use as
electrodes for the remelting process. The VAR process improves
the purity and uniformity of the metal being produced.
The solid steel electrode is lowered on a control rod into a ver-
tical vacuum chamber. The control rod acts as a cathode (positive
terminal) when an electric current is passed through it into the
steel. The bottom half of the vacuum chamber is actually a water-
cooled mold, which serves as the anode (negative terminal). When
the current is turned on, the solid steel performs like a giant elec-
trode in arc welding, with the heat of the electric arc melting the
end of the steel electrode.
During the remelting process, gaseous impurities are drawn
off by the vacuum in the chamber as the molten steel drops into
the water-cooled mold below. Once solidified, remelted product is
almost free of center porosity and gaseous inclusions are greatly

17

Ch02.pmd 17 10/29/02, 11:12 AM


Tool and Die Making Troubleshooter

reduced. From the standpoint of cleanliness, VAR steels are equiva-


lent to ESR steels. However, the ESR process has the ability to
produce lower sulfur contents.

Vacuum Induction Melting


The VIM process melts and refines steel in an induction fur-
nace situated in a vacuum chamber. Scrap—or, in some instances,
molten steel conventionally melted in an electric arc furnace—is
charged into the furnace and melted. During this process, the un-
desirable gases are removed by the vacuum pump. Once the gases
are eliminated, the furnace tilts and pours the refined steel into a
trough to convey it into a holding ladle in an adjoining vacuum
chamber. This holding chamber is positioned so that molds can be
rolled underneath it. All operations are remotely controlled from
a panel adjoining the vacuum chambers.
VIM can purify steels by removing dissolved gases and volatile
contaminants during the process. This is achieved in three ways:
(1) the evolution of gases dissolved in the metal, (2) the removal
of impurities by chemical reactions, and (3) the vaporization of
impurities with higher vapor pressures than that of the base metal.
The net result is very clean steel, with a very low amount of non-
metallic inclusions.

Single Melt Versus Special Melting/Refining


Figures 2-7 and 2-8 show the type of metallurgical improve-
ment achieved through the use of special refining and/or remelt-
ing techniques. Identical cross sections of a 4-in. (10.2-cm) thick
AISI-S7 tool-steel plate are shown. Each sample was cut from di-
rect center locations.
Figure 2-7 is a sample of conventional single-melt material. Fig-
ure 2-8 is a sample of ESR quality. Note the absence of inclusions
in the ESR material. The cleaner, more uniform, sounder ESR
material will have significantly improved mechanical properties
due to the reduced inclusion levels.

18

Ch02.pmd 18 10/29/02, 11:12 AM


Chapter 2: Steelmaking Methods and Hot Working

Figure 2-7. Hot-acid-etched, conventional single-melt tool steel. Note the


centerline inclusions and porosity.

Figure 2-8. Hot-acid-etched, ESR remelted tool steel. Note the absence of
centerline inclusions and porosity.

POWDER-METAL TOOL STEELS


Powder metallurgy has become an increasingly important pro-
cess for producing high-speed steels, high-temperature alloys, and

19

Ch02.pmd 19 10/29/02, 11:12 AM


Tool and Die Making Troubleshooter

unique steels with unusually high alloy content and/or element


combinations.
Powder metallurgy produced tool steels are made by compact-
ing metal powders under conditions of high temperature and pres-
sure, a process called hot-isostatic pressing (HIP). The “compacts”
produced are then rolled or forged to produce desired bar sizes,
which are then handled as conventional tool-steel bars (annealed
and ground before shipment). The ultimate user will process the
powder metallurgy produced steel in the same manner as conven-
tionally produced tool steel—by machining and heat treating.
Metal powders are made by atomizing molten metal or alloys
(Figure 2-9).
The molten material is poured through a small nozzle and an
atomizing gas produces a rapidly solidified fine powder. The pow-
der is then screened to select the powder size of choice. Metal
powders of iron and various alloys may then be blended to pro-
duce a desired chemical composition.
In one producer’s proprietary process, the chemical analysis of
the steel is achieved via conventional steelmaking. Then, instead

Figure 2-9. Atomizing molten metal or alloys to produce a rapidly solidified


metal powder.

20

Ch02.pmd 20 10/29/02, 11:12 AM


Chapter 2: Steelmaking Methods and Hot Working

of teeming the steel into ingot molds as in conventional steelmak-


ing, the steel is gas atomized to produce the metal powder that is
later compacted under heat and pressure to produce a “compact”
or ingot for further processing. However, most producers take
screened metal powders, either blended or produced by this pro-
cess, and pour them into a steel container that is evacuated and
sealed. The container is then subjected to high temperature (usu-
ally the forging temperature for the steel) and pressure (either
high-pressure gas or mechanical) to produce the compact. The
compact is then subjected to rolling or forging and further pro-
cessing as if it were an ingot produced by conventional processing.
The powder metallurgy process is expensive when compared to
conventional single melting, VAR, and ESR. However, it has the
ultimate advantage of producing steels free of the effects of chemi-
cal segregation and large, poorly distributed carbides. The net
result of the material’s lack of chemical segregation, smaller car-
bide size, and more uniform carbide distribution is finer grain-
size control and a finished tool with more uniform properties after
heat treatment. The process also allows production of steels with
finer, more uniformly distributed sulfides. These steels can be
resulfurized to higher sulfur levels for enhanced machinability
while maintaining high toughness properties. This cannot be ac-
complished with conventional steel production processes. Because
of the need for small batch processing, the powder metallurgy
process is applicable primarily to the more expensive grades of
tooling steels, high-speed speeds, and superalloys (Crucible Mate-
rials Corp. 1995; United States Steel Corp. 1971).

HOT WORKING OF TOOL STEELS


The hot working of tooling steels, whether by forging and/or
rolling, is performed to reduce the cross-sectional area, improve
the internal structure, and shape the metal into desired bars,
sheets, or plates.
After the steel has been melted in a conventional steelmaking
process, carefully refined, and alloyed, it is tapped from the melt-
ing furnace into a ladle and then teemed (poured) into molds to
form ingots. After the molten ingots solidify, they are stripped

21

Ch02.pmd 21 10/29/02, 11:12 AM


Tool and Die Making Troubleshooter

from their molds and reheated to the proper hot-working tem-


perature for forging (often referred to as “cogging”) or rolling.
The ingots are then forged and/or rolled (usually after an addi-
tional heating operation) into the finished product.
Forging breaks up the coarse crystalline structure of the ingot,
redistributes segregation, compacts any sponginess in the ingot
center, welds minor defects, and reduces porosity. In addition, it

Figure 2-10. Hammering was the first method of forging (Lincoln Electric
Co. 1980).

22

Ch02.pmd 22 10/29/02, 11:12 AM


Chapter 2: Steelmaking Methods and Hot Working

generally reduces the size of massive nonmetallic inclusions and


breaks them up.
As shown in Figure 2-10, hammering was the first method of
forging. It evolved from manually striking hot metal with a heavy
hammer to the use of large power hammers driven by gravity and
steam.
With the advent of high-powered forging machines, larger pieces
of hot metal could be reduced between a top die and bottom anvil.
As forging techniques further evolved, steam and gravity ham-
mers gave way to pneumatic and hydraulic presses that squeeze
and knead the hot workpiece to produce deeper hot working than
can be obtained by hammer forging or rolling.
During rolling, metal is forced to pass between two rolls spin-
ning in opposite directions (Figure 2-11). As the metal passes be-
tween the rolls, it is reduced in cross-sectional area and elongated
longitudinally. This also breaks up nonmetallic inclusions and elon-
gates them. Rolling refines the structure of the steel, making it
more fine-grained, uniform, homogeneous, and dense.

Figure 2-11. Reduction and elongation during rolling or forging (Forging


Industry Association 1993).

23

Ch02.pmd 23 10/29/02, 11:13 AM


Tool and Die Making Troubleshooter

Compared to rolling, forces from forging penetrate more deeply


into the workpiece and the forces from press forging penetrate more
deeply than with hammer forging. Therefore, the forging process
more effectively breaks up impurities and refines and improves the
resulting ingot products (such as billets, bars, and plates). This is
why forging is preferred for high-alloy steels.
The amount of reduction required to produce a sound product
varies with the type of ingot and steelmaking process used. Most
tool steels produced today by conventional melting methods (basic
oxygen furnace, electric furnace, argon-oxygen decarburization)
receive a minimum of 3:1 or 4:1 reduction (based on the change in
cross-sectional area) during rolling or direct forging from ingot to
final product size. The process may incorporate more than one
heating operation during the hot working. Consumable-electrode
product, such as ESR or VAR, may be rolled or direct forged with a
lesser amount of reduction (2:1), because such ingots are more sound
than those produced via conventional single-melting methods.

CAST-TO-SHAPE VERSUS WROUGHT TOOL STEELS


The decision to produce tools and dies from castings (cast to
shape) is usually cost driven. By comparison, if made from wrought
hot-worked steel (that is, cast and then forged and/or rolled), these
same tools and dies would require more expensive machining.
However, in making the decision to go to a cast-to-shape product,
the tooling designer must understand some basic metallurgical
quality and physical property differences between the two pro-
cesses.
Hot working affects the strength, directional properties, inter-
nal defects, toughness, ductility, and heat treatment of steel. Cast-
ing by itself will not obtain the strengthening effects imparted
from hot working (casting and forging and/or rolling). A casting
has neither grain flow nor directional strength, two properties
imparted from hot working after a material has been cast. Pin-
holes, alloy segregation, dendritic structures, and other forms of
unsoundness and imperfections are compressed, elongated, dis-
persed, and thus refined by the hot-working process. The grain-
refining effects of hot working reduce brittleness and increase

24

Ch02.pmd 24 10/29/02, 11:13 AM


Chapter 2: Steelmaking Methods and Hot Working

impact strength, toughness, and ductility. Uniform, homogeneous


structures developed and enhanced by hot working respond bet-
ter to heat treatment than as-cast and/or cast-to-shape structures.
Note that there are many cost, technical, and metallurgical dif-
ferences between cast and hot-worked tool steels. It is always in
the best interest of the end user to carefully identify when these
two very different processes are most appropriate in terms of con-
struction costs, techniques, and production application (Forging
Industry Association 1994–95).

REFERENCES
Allen, Dell K. 1969. Metallurgy Theory and Practice. Homewood,
IL: American Technical Publishers, Inc., p. 574.
American Society for Metals. 1948. Metals Handbook. Cleveland,
OH: American Society for Metals, pp. 325–329.
Bethlehem Steel Corp. 1974. Tool Steel Topics. Issue 211, Sept./
Oct. Bethlehem, PA: Bethlehem Steel Corp.
Crucible Materials Corp. 1995. Crucible Particle Metallurgy.
Oakdale, PA: Crucible Materials Corp.
Forging Industry Association. 1993. Open Die Forging Technol-
ogy. Cleveland, OH: Forging Industry Association, p. 7.
——. 1994–95. Custom Forging Capability Guide. Cleveland, OH:
Forging Industry Association, pp. 2–3.
Lincoln Electric Co. 1980. Metals and How to Weld Them, 2nd Edi-
tion. Cleveland, OH: Lincoln Electric Co., p. 2.
Shelby, Oleg D., and Wadsworth, Jeffrey. 1985. “Damascus Steels.”
Scientific American, February, p. 112.
Stoughton, Bradley. 1934. The Metallurgy of Iron and Steels. New
York: McGraw-Hill Co., Inc., p. 2.
Sunday Democrat and Chronicle. 1981. “Secret of Damascus Un-
locked, Metallurgists Simulate Steel in Legendary Swords.” Oc-
tober 4, p. 8D. Rochester, NY: Gannett Co., Inc.

25

Ch02.pmd 25 10/29/02, 11:13 AM


Tool and Die Making Troubleshooter

United States Steel Corp. 1971. The Making, Shaping and Treat-
ing of Steel, 9th Edition. Pittsburgh, PA: United States Steel Corp.,
p. 403 and p. 551.
——. 1985. The Making, Shaping and Treating of Steel, Tenth
Edition. Pittsburgh, PA: United States Steel Corp., pp. 479–689.

26

Ch02.pmd 26 10/29/02, 11:13 AM


Chapter 3: Quality Considerations

3
Quality Considerations

Guaranteeing consistent results in any manufacturing operation


today requires establishing a comprehensive quality system that
covers all of the standards, processes, and procedures used to
produce a product or service.
In recent years, the International Organization for Standard-
ization’s ISO 9000 quality system has evolved into the leading
global quality standard. This chapter examines the importance
and application of ISO 9000 standards and methodology to air-
craft-quality and mold-quality steels.

ISO 9000 QUALITY SYSTEMS


In today’s competitive environment, companies have realized
the advantage of developing, implementing, and managing qual-
ity systems. Quality systems help companies organize efforts to
meet customer requirements at the lowest possible cost. Many
manufacturers have proven that a good quality system pays ma-
jor dividends over and above the total cost of training personnel
in the quality standards and the investment in implementing and
maintaining the processes and procedures.

Say What You Do, Do What You Say


The best explanation of why the ISO 9000 quality system has
become the most widely implemented global quality standard is
because, in its simplest form, it defines quality goals and then

27

Ch03.pmd 27 10/29/02, 11:13 AM


Tool and Die Making Troubleshooter

Figure 3-1. The quality triangle (Stat-a-Matrix 1993).

confirms that they are being met. “Say what you do and do what
you say,” and good results will follow. It is a system for enforcing
continuous improvement.
The ISO standard incorporates four levels of organization for
compliance (Figure 3-1). Level 1 requires writing a quality manual
that defines the company’s quality policy and assigns responsibility
to particular job areas. Level 2 requires writing quality-system
procedures that define specifically who does what to ensure that
everyone understands all facets of the quality policy and maintains
compliance with it. Level 3 requires writing detailed work
instructions that carefully define exactly how each job must be
done. Finally, Level 4 requires keeping records and forms on file

28

Ch03.pmd 28 10/29/02, 11:13 AM


Chapter 3: Quality Considerations

to provide evidence that all of the required quality procedures


have been met.
As the ISO quality system grew and evolved, the Big Three auto
manufacturers worked with the American Society for Quality
Control (ASQC) to adopt a common quality standard (QS) specifically
for their industry. This standard soon became known as QS 9000.
Working within the framework of ISO 9000, the automotive-
driven quality standard added a set of industry-specific requirements,
such as expanded quality planning, more stringent design controls,
system analysis, record retention, control of reworked product,
etc. The QS 9000 quality system incorporates ISO 9000 in its
entirety. Hence, manufacturers certified compliant to QS 9000 are
also in compliance with other ISO-based quality approaches.
Recently, a tooling and equipment (TE) supplement has been
added to the QS 9000 standard. It is designed to assist toolrooms
and machine shops in developing a quality program that focuses
on and manages their unique concerns and requirements. When
properly implemented, the guidelines in the TE supplement lead
to standardization, consistency, reduced costs, and improved rela-
tionships between suppliers and original equipment manufac-
turers (OEMs).

Quality for the Bottom Line


Quality systems and standards—as they apply to ISO 9000, QS
9000, and the TE supplement—combine to provide much more than
quality improvements in products produced. A study of business
performance in more than 200 mechanical manufacturing com-
panies before and after ISO 9000 certification showed an increase
of 6.8% in profits in smaller firms (fewer than 200 employees)
and a 4% increase in profits in larger firms.
In summary, the organization and implementation of continual
compliance to strict quality system requirements will (Stat-a-
Matrix 1993):
• Improve the effectiveness of a company’s management system.
• Control the quality of supplied products.
• Streamline handling, storage, packaging, and delivery systems.
• Ensure the accuracy of inspection, measuring, and test equip-
ment.

29

Ch03.pmd 29 10/29/02, 11:13 AM


Tool and Die Making Troubleshooter

• Control nonconforming product.


• Take corrective and preventive action against quality concerns.
• Implement an effective continuous-improvement process.
• Increase profitability.

AIRCRAFT-QUALITY STEEL
Aircraft-quality steel denotes steels intended for important or
highly stressed parts and components in the aerospace industry.
These steels are also used in other applications where their
particular quality levels are deemed necessary for an intended
application, whether aerospace or not. Special steelmaking
practices, more rigid inspection techniques, and more restrictive
selection are necessary to meet the aerospace industry’s rigid
standards.

Two Quality Levels


Two levels of quality have been accepted by producers and users
of aircraft steels, both based on meeting particular magnetic-
particle requirements. Both specifications detail the specimen
preparation, the magnetic-particle inspection procedure to be used,
rating of the magnetic-particle indications, and acceptance
standards.
Level 1, aircraft quality per specification AMS 2301 (SAE
International 2001B). This quality classification can be met
consistently with air-melted, vacuum-degassed steel heats. It can
also be met with some air-melted steel heats that are not vacuum
degassed.
Level 2, premium aircraft quality per specification AMS
2300 (SAE International 2001A). This quality classification is
meant for the most critical parts and components, has restrictive
magnetic-particle acceptance standards, and requires the use of
consumable-electrode remelted steels—electroslag remelting
(ESR), vacuum arc remelting (VAR), etc.

30

Ch03.pmd 30 10/29/02, 11:13 AM


Chapter 3: Quality Considerations

Testing for Aircraft Quality


Under certain circumstances, a purchaser of SAE 4340 steel
may wish to have the steel tested to see if it will meet certain
minimum transverse mechanical-property requirements (trans-
verse strength and ductility levels). Tests are performed on heat-
treated transverse tensile specimens secured from representative
locations in a heat of steel.
The testing procedure includes the machining of oversize trans-
verse tensile specimens from the mid-radius location of billet
material, heat treating the oversize specimens to ultra-high ten-
sile and yield strengths, final machining of the tensile specimens,
and tensile testing. Two specifications apply:
• AMS 6415 (SAE International 1997)—steel produced to this
specification must also meet the magnetic-particle require-
ments of AMS 2301. The tensile requirements can be met
with air-melted steels, vacuum degassed or not.
• AMS 6414 (SAE International 1998)—steel produced to this
specification must also meet the magnetic-particle require-
ments of AMS 2300. The tensile requirements require, as do
the magnetic-particle requirements of AMS 2300, consum-
able-electrode remelted steels such as VAR and ESR.

MOLD-QUALITY STEEL
The designations “superior mold quality,” “mold quality,” and
“plastic mold quality” are frequently used to identify tool steels
specifically made for applications where tooling costs, steel quality,
reliability, and production performance are of paramount
importance. Mold-quality steels are specifically formulated and
manufactured to be cleaner and more homogeneous (uniform) than
conventionally manufactured tooling steels. Originally, these
designations were applied to steels used primarily by plastic
injection-mold builders. Today’s mold-quality steels have become
increasingly popular for more general tooling applications as tool
builders look for steels with better metallurgical properties. The

31

Ch03.pmd 31 10/29/02, 11:13 AM


Tool and Die Making Troubleshooter

latter manifest increased toughness (higher impact strength),


better ductility (higher elongation and reduction of area), and
improved polishing characteristics. All of these are characteristics
of cleaner and more homogeneous steels.
Manufacturers of mold-quality steels use processes that incor-
porate electric-furnace melting and/or argon-oxygen decarburiza-
tion (AOD) refining along with vacuum degassing. This is usually
followed by a second melting process such as ESR, VAR, or vacuum
induction melting (VIM). In combination, these processes work
to reduce, filter out, and disperse inclusions and porosity, and en-
sure that any remaining defects in the steel are small. However,
even when superior manufacturing processes such as ESR, VAR,
and VIM are used and followed by careful inspection, a mold-qual-
ity steel will never be totally free of inclusions and porosity.
There are many ways to check for the presence of nonmetallic
inclusions in steel. Macroscopic examination, macroetching, and
magnetic-particle testing are a few of the more popular methods.
However, microscopic methods are frequently used to more accu-
rately determine the size, number, distribution, and type of inclu-
sions present.
The nonmetallic inclusions are rated microscopically in accor-
dance with specification ASTM E-45 (JK microcleanliness test-
ing), which classifies them by type, size, and frequency. Rating of
inclusion types and size is usually referred to as “qualitative
microcleanliness rating,” while the counting of the number of in-
clusions of each rated size is referred to as “quantitative
microcleanliness rating.” The inclusion categories are Type A (sul-
fides), Type B (aluminates), Type C (silicates), and Type D (globu-
lar oxides). Inclusion ratings are shown in Figure 3-2.
Although tool and mold steels are generally not sold to specific
maximum microcleanliness standards, Figure 3-2 provides ap-
proximate information on expected cleanliness levels that can be
obtained with various melting processes. These rating values are
typical and may vary with individual producing mill capabilities,
practices, and standards for quality.
The JK microcleanliness rating of inclusions is conducted mi-
croscopically on polished microsamples (taken from the mid-ra-
dius of billets) at a magnification of 100×. The inclusions viewed

32

Ch03.pmd 32 10/29/02, 11:13 AM


Chapter 3: Quality Considerations

ASTM E 45, Method “D “D”” JK Rating


ESR and V acuum-remelted Steels
Vacuum-remelted
These steels are first melted in an electric furnace, vacuum degassed and/or
AOD refined, and then remelted in an ESR or vacuum-type furnace.

Type A Type B Type C Type D


Thin Thick Thin Thick Thin Thick Thin Thick
1.0 0 1.0 0 1.0 0 1.0 0

JK Rating
Electric FFurnace,
urnace, V acuum Degassed, and/or A
Vacuum OD
AOD -refined Steels
OD-refined
These steels are electric-furnace melted, then refined via vacuum degassing
and/or AOD processing. They are not remelted.

Type A Type B Type C Type D


Thin Thick Thin Thick Thin Thick Thin Thick
2.5 1.5 2.0 1.0 0.5 0.5 1.0 1.0

JK Rating
Electric-furnace Melted Steel
No vacuum degassing or AOD refining is performed

Type A Type B Type C Type D


Thin Thick Thin Thick Thin Thick Thin Thick
2.5 1.5 2.0 1.5 2.0 1.5 2.0 1.5

Figure 3-2. Inclusion ratings for various mold-quality remelted steels.

are compared to those shown on a separate chart for each type in


ASTM E-45 (ASTM International 1997) to identify maximum size,
both thin and thick, for the A, B, C, and D-type inclusions, respec-
tively.
An inclusion is placed in the “thin” category when its width is
up to 5 ␮in. (0.13 ␮m), and in the thick category when its width
is over 5 ␮in. (0.13 ␮m). The length of an inclusion for each cat-
egory is also determined with an applicable chart from ASTM E-
45. An inclusion rating of 0.5 represents a maximum inclusion
length of 0.15 in. (3.8 mm) at 100× magnification. A 1.0 rating
represents an inclusion length of 0.30 in. (7.6 mm) at 100× mag-
nification, and a 1.5 inclusion rating represents a length of 0.70
in. (17.8 mm).

33

Ch03.pmd 33 10/29/02, 11:13 AM


Tool and Die Making Troubleshooter

REFERENCES
ASTM International. 1997. “Standard Practice for Determining
the Inclusion Content of Steel.” ASTM E-45. West Conshohocken,
PA: ASTM International.
SAE International. 2001A. “Premium Aircraft-quality Steel Clean-
liness Magnetic Particle Inspection Procedure.” Aerospace mate-
rial specification (AMS) 2300, Rev. J. Warrendale, PA: Society of
Automotive Engineers.
——. 2001B. “Cleanliness, Aircraft-quality Steel Magnetic Par-
ticle Inspection Procedure.” Aerospace material specification
(AMS) 2301, Rev. J. Warrendale, PA: Society of Automotive Engi-
neers.
——. 1998. “Steel, Bars, Forgings, and Tubing 0.80 Cr-1.8 Ni-0.25
Mo (0.38-0.43 C) (SAE 4340) Vacuum Consumable-electrode Re-
melted.” Aerospace material specification (AMS) 6414 Rev. H.
Warrendale, PA: Society of Automotive Engineers.
——. 1997. “Steel, Bars, Forgings, and Tubing, 0.80 Cr-1.8 Ni-
0.25 Mo (0.38-0.43 C) (SAE 4340).” Aerospace material specifica-
tion (AMS) 6415, Rev. P. Warrendale, PA: Society of Automotive
Engineers.
Stat-a-Matrix. 1993. “ISO 9000 Seminars—Lead Auditor (Asses-
sor) Training.” Edison, NJ: Stat-a-Matrix, pp. 2-1 to 2-43, and 3-2.

34

Ch03.pmd 34 10/29/02, 11:13 AM


Chapter 4: Metallurgy and Engineering Considerations

4
Metallurgy and Engineering
Considerations

This chapter examines the basic metallurgical and physical char-


acteristics of steel that are particularly important for tooling steels.

BASIC MECHANICAL PROPERTIES


The basic strength properties of a metal are tensile, impact,
compressive, fatigue, torsion, and shear strengths.
Tensile strength is the resistance of a material to a force acting
to pull it apart, a key property in evaluating metals. Metallurgists
measure the tensile strength of a metal by determining the maxi-
mum load in pounds per square inch (psi [kPa]) when breakage
occurs. This maximum load is also known as ultimate tensile
strength.
Stress is the amount of force a load applies to a material, either
in tension or compression.
Strain is the physical effect of a stress. In the case of tensile
stress, the resulting strain in the material is measured by the
amount of stretching or elongation that takes place.
Yield point is the minimum stress at which an increase in strain
occurs without an increase in stress. This term is strictly appli-
cable to mild steels, where the yield point is considered to be its
yield strength. For other steels, yield strength is the stress re-
quired to strain the test piece by a specified small amount beyond
the elastic limit.

35

Ch04.pmd 35 10/29/02, 11:14 AM


Tool and Die Making Troubleshooter

Elastic limit is the maximum stress the metal will support with-
out permanent deformation (American Society for Metals 1948).
Any stress below this point will stretch the metal, but when the
force is removed the material will elastically return to its original
dimensions, like a rubber band. For commercial purposes, the yield
strength is considered to be identical to the elastic limit.
A metal with a high yield point or high elastic limit is needed
for parts that must hold their exact shape and size. A low yield
strength is an advantage when a part is to be cold bent or formed
because less force is required to form it.
The modulus of elasticity is the ratio of stress to strain. It is
used to compare the stiffness of one metal to another. This can be
determined from the slope of the stress-strain curve in tensile
testing, within the elastic limit. A material that stretches easily
has a low modulus and a low slope of the stress-strain curve; that
is, a low stress divided by a high strain. In a tougher material
where a high stress produces a small strain, the modulus is a higher
figure.
For example, steel is two to three times stiffer than cast iron. If
two bars of equal size—one cast iron and one steel—are stressed
in tension, the cast-iron bar will deflect twice as far as the steel
bar. If two steel bars are similarly stressed, even though one may
be hard tool steel and the other soft mild steel, they will deflect
the same amount.
Although steel is stiffer than cast iron, it is more ductile. The
cast-iron bar will break suddenly with a brittle fracture as the
stress is increased, while the steel will continue to deform (stretch
and “neck down”) prior to failure as the stress is increased. The
modulus of elasticity is not a measure of the amount of stretch a
particular metal can take before breaking or deforming. It simply
tells how much stress is required to make the metal stretch a
given amount.
Ductility is the ability of a metal to stretch and become perma-
nently deformed without breaking or cracking. Ductility is mea-
sured by the percentage of reduction in cross-sectional area and
the percentage of elongation of the test bar. Ductile steels are nec-
essary for formed and drawn steel parts such as automobile body
panels and fenders. A metal with high ductility will stretch before

36

Ch04.pmd 36 10/29/02, 11:14 AM


Chapter 4: Metallurgy and Engineering Considerations

breaking. A metal with low ductility fails suddenly with a brittle


fracture.
Brittleness is the opposite of ductility. Brittle materials show
practically no permanent distortion before failure and often fail
suddenly without warning. Glass is an excellent example of this.
Ductility and elasticity must not be confused. Ductility is the
ability to permanently elongate, when stretched, without break-
ing. Elasticity is the ability to elongate and return to original size.
All ductile metals have considerable elasticity because they will
return to the original length if they are not stretched beyond a
certain point. A metal may be elastic, however, without being duc-
tile. It stretches up to a certain point at which it suddenly breaks,
instead of stretching further as a ductile metal would do.
Cast iron’s modulus of elasticity is lower than that of steel, and
it stretches more easily, but not very far since it has a low elastic
limit. It breaks when the load reaches the elastic limit. Steel can
withstand higher stresses than cast iron and still return to its
original shape.
The compressive strength of a material is its ability to resist
compressive loads. Rubber, for example, has a low compressive
strength, while steel has a high compressive strength and can carry
extremely heavy loads without deformation. The relationship of
compressive strength to tensile strength varies according to the
material group. In the case of most steels and the aluminum and
magnesium alloys, these values are approximately equal. Low-
strength cast irons may have compressive strength values several
times their tensile strengths.
Fatigue strength is a measure of a material’s ability to resist
continual reversals in stress. A railroad-car axle is stressed in ten-
sion at the top (along its axis) and in compression at the bottom.
At any point on the axle, each revolution produces a change from
tension to compression. Metals will fail under this type of chang-
ing load due to fatigue failures that usually start with a small
crack that becomes progressively larger under the repeated stress
cycle until failure occurs. Fatigue strength is the ability to resist
fracture when the stresses are variable and alternate through a
cycle. Fatigue tests are made by subjecting a test specimen to vari-
able loads, alternately bending, rotating, or tensioning it.

37

Ch04.pmd 37 10/29/02, 11:14 AM


Tool and Die Making Troubleshooter

Fatigue limit, or endurance limit, is the maximum stress that a


metal can withstand for a predetermined number of repetitive
cycles. For steel, the endurance limit varies between 40–60% of
the ultimate strength of the material in tension. The exact limit
depends both on the nature of the material and its surface condi-
tion.
Fatigue failures start at the surface as tiny cracks, which spread
into the metal until failure occurs. Scratches, tool marks, threads,
and other surface irregularities act as places for cracks to initiate.
The fatigue strength of a specimen with lathe-tool marks may be
half that of one with a smooth, polished surface. Sharp changes in
the cross section of a piece of metal, such as a shaft shoulder or a
keyway, also reduce fatigue strength. Producing welds on metals
that must have a high fatigue strength is an exacting job. Besides
matching the strength of the parent metal, the weld must not be
the source of any undue hardening or softening of the parent metal.
Hardness is an important property. Once the hardness of a metal
is determined, some of its other properties can be estimated. The
tensile strength of steel, for example, increases directly with an
increase in Brinell hardness.
Toughness is a term often used rather loosely. In many instances,
it relates to the combination of properties that enables a metal to
stand up against certain stresses in specific applications. Usually,
toughness means the general ability of a metal to withstand the
shock of a rapidly applied load. Some common metals ranked ac-
cording to their relative toughness (most to least tough) are: cop-
per, nickel, iron, magnesium, zinc, aluminum, lead, tin, and cobalt.
Notch effect describes the effect a surface-stress concentrator
has on the toughness of a material. Such conditions include deep
V-grooves, cracks, tool marks, and sharp inside corners or other
sudden changes in a part’s cross section. Some steels, like high-
alloy tool steels, are especially sensitive to the notch effect. Oth-
ers have a high degree of notch toughness, meaning their toughness
remains relatively high in the presence of notches or similar con-
ditions.
Notch-sensitive metals fail more readily than others under im-
pact or repetitive loading if a notch is present. The notch sensitiv-
ity of a metal usually increases as the hardness of the metal
increases and as the operating temperature decreases. In general,

38

Ch04.pmd 38 10/29/02, 11:14 AM


Chapter 4: Metallurgy and Engineering Considerations

a metal that shows extreme strength under impact as a notched


specimen will demonstrate great toughness under most service
conditions.

Toughness versus Brittleness


Toughness and brittleness are often thought of as direct oppo-
sites. This is true to a great extent and is helpful in the study of
failures under actual service conditions.
In a tough fracture, the two pieces of metal look like they were
torn apart. The broken surfaces are very irregular and, in many in-
stances, show the fiber-like nature of the metal’s internal structure.
These surfaces are usually dull in appearance. A tough fracture is
often called a ductile fracture, especially when the metal has high
ductility and there is considerable reduction in the cross section
and a corresponding elongation in the area of fracture.
In contrast, a brittle fracture frequently looks like the two pieces
were sheared apart. In some cases, the metal shatters into mul-
tiple pieces or many small fragments. The broken surfaces are
clean, sometimes smooth, but often with jagged edges. The sur-
faces usually show the crystal structure of the metal rather than
a fibrous-like quality, and generally are bright in appearance.
Although metals may be identified as tough or brittle on the
basis of their appearance when fractured at ambient temperatures,
some caution is needed. A metal that appears to have great tough-
ness and produces a tough or ductile fracture at ambient tem-
perature may fail under lighter loads and produce a brittle fracture
at a lower temperature. Metallurgists call the point at which this
change from ductile to brittle fracture occurs the transition tem-
perature (Lincoln Electric 1973; Bethlehem Steel Corp. 1980).

HYDROGEN EMBRITTLEMENT
Hydrogen embrittlement is a condition of low ductility in metal
that results from the absorption of hydrogen gas. Steels frequently
absorb hydrogen during the manufacturing operation or subse-
quent operations such as acid-bath pickling, chrome plating, or
welding.

39

Ch04.pmd 39 10/29/02, 11:14 AM


Tool and Die Making Troubleshooter

The amount of hydrogen absorbed by a steel during its melting


phase depends on its chemical analysis, atmospheric moisture,
and other factors. Thermal processing, such as normalizing and
annealing, is used for steels that contain hydrogen. These methods
diffuse some of the gas throughout the metal. However, not all of
the hydrogen diffuses and some of it causes internal pressures to
develop, particularly in and around nonmetallic inclusions.
Gaseous hydrogen can also precipitate in microvoids, resulting
in high-pressure pockets of gas. The pressure may be high enough
to form discontinuities called flakes that frequently propagate in
service, causing hairline cracks and/or brittle failures. Manufac-
turers of tool steels are very careful to vacuum degas molten steels
to remove excess hydrogen and improve internal cleanliness.
Although tool steels are not particularly susceptible to prob-
lems related to hydrogen flaking, tools and dies can absorb hydro-
gen from operations such as chrome plating, welding, and pickling,
and during exposure to certain service conditions such as hydro-
gen-sulfide atmospheres in “sour” oil-field environments. Hydro-
gen absorption in hardened steels with martensitic structures
causes a most severe type of embrittlement. Absorbed hydrogen
can also result in embrittlement and cracking when stresses are
applied.
Hydrogen absorption (embrittlement) from chrome plating,
pickling, and welding can be diffused from the steel by baking at a
temperature range of at least 350–400° F (177–204° C). After heat-
ing to the baking temperature, the steel should be soaked at that
heat level for approximately 2 hours. A slow cooling should follow.
Postheating following welding is also a good procedure for diffus-
ing hydrogen (United States Steel 1971).

ALLOY SEGREGATION (BANDING)


Chemical segregation, or banding, occurs in some tool steels
when alloy-rich bands precipitate as the molten steel solidifies in
the ingot mold. This phenomenon is most characteristic of the
chromium-molybdenum steels (S7, A2, D2, and M2). The bands
of chemical segregation, when observed under magnification are
oriented longitudinally (Figure 4-1). They become more preva-
lent as the distance from the bar surface to the interior increases.

40

Ch04.pmd 40 10/29/02, 11:14 AM


Chapter 4: Metallurgy and Engineering Considerations

Figure 4-1. Photomicrograph of “banded” D2 tool steel (50x).

The photomicrograph (50× magnification) in Figure 4-1 shows


a banded condition on a D2 tool steel sample (in the annealed
condition) that has been cross-sectioned, polished, and etched. The
light areas are carbides. Note the two segregated, alloy-rich car-
bide bands running longitudinally.
Banding is more likely to occur in conventionally produced steel
ingots because of how they solidify; that is, from the outside of
the ingot inward. Banding has also been observed, although to a
lesser degree, in electroslag remelt (ESR) and vacuum arc remelt
(VAR) ingots—where solidification takes place along the vertical
axis. No steelmaking process to date has been able to eliminate
the occurrence of banding in steels prone to this phenomenon.
Alloy segregation and banding are not a cause of concern for
most tool-and-die applications. It is at times a problem on highly
polished, optical-quality plastic-injection-mold cavities because the
bands generally cannot be polished out of the finished surface.
However, alloy segregation or banding also may be a problem for
tooling that must be hardened to very uniform, finite tolerances

41

Ch04.pmd 41 10/29/02, 11:14 AM


Tool and Die Making Troubleshooter

throughout a longitudinal cross section. It is often possible to cir-


cumvent this problem by selecting stock sizes that will allow criti-
cal surfaces to be located away from center locations where banded
areas may be concentrated.

THE ART OF SPARK TESTING


When a high-speed grinding wheel is touched to a piece of steel,
small chips of metal are torn away and the heat generated is so
great that it causes the chips to burn in the form of bright sparks
(Figure 4-2). Each different grade of steel will burn with a unique
spark burst, pattern, shape, and color depending on its chemical
composition. By carefully comparing the spark burst, intensity,
colors, and spark shapes with known samples, an experienced
spark tester can identify the steel by its grade and chemistry with
surprising accuracy.
The best way to learn how to do spark testing is to gather
samples that have been carefully identified with regard to grade
and hardness, and to practice using the following guidelines. Keep
in mind that each individual spark tester may make slightly dif-
ferent observations depending on the spark environment, the sur-
rounding light, and his/her eyesight. Practice daily with known
samples until their sparks are recognized. Do not use a brightly
lit room or one that is too dark. Diffused lighting is best.
Some tips for learning the art of spark testing:
• Use a high-speed (10,000–20,000 rpm) grinder.
• Use an abrasive wheel 2–3 in. (50.8–76.2 mm) in diameter
and approximately 40 grit.

Figure 4-2. A typical spark-burst pattern (Carpenter Technology 1967).

42

Ch04.pmd 42 10/29/02, 11:14 AM


Chapter 4: Metallurgy and Engineering Considerations

• Dress the wheel so it is not loaded with grit from previous


grinding.
• Avoid using pressure that slows the abrasive wheel.
• Do not spark test on a decarburized and/or scaled surface layer.
• The higher the carbon level, the greater the spark-burst in-
tensity.
• Hard steels throw a longer, more dense pattern than annealed
steels.
Alloy additions suppress bursts while adding color and shape to
the sparks. Each alloy element burns with a unique spark shape
and color (Bethlehem Steel Corp. 1972; Carpenter Technology 1967):
• The more carbon, the more plentiful and complicated the
burst.
• Silicon suppresses the carbon burst and burns dark red.
• Chromium suppresses the spark stream and burns bright
orange.
• Nickel suppresses the spark stream and burst slightly and
causes the spark stream to end with a forked tongue.
• Tungsten suppresses the effects of all other elements in the
spark stream and burns with a bright orange tongue pattern.
• Vanadium tends to brighten the spark stream as a whole.
• Molybdenum causes a characteristic spear point.
• Manganese has a minimal effect on the spark burst of steels
containing moderate or large amounts of other alloys.

EFFECTS OF ALLOYING ELEMENTS


It is important to understand the effects of alloying elements in
steels (Bethlehem Steel Corp. 1972, 1980; Timken Steel 1986):
• Aluminum (Al) is used as a deoxidizer. It controls grain
growth, promoting fine-grained steels. It is used in nitriding
steels because of its strong tendency to form aluminum ni-
trides, which contribute to wear resistance.
• Boron (B) is used to increase the hardenability of carbon and
alloy steels. Its greatest effect is in low-carbon varieties of both.
• Carbon (C) is the principle hardening element in steel. It in-
creases hardness, tensile strength, and wear resistance, but
reduces ductility.

43

Ch04.pmd 43 10/29/02, 11:14 AM


Tool and Die Making Troubleshooter

• Chromium (Cr) improves hardenability, wear, corrosion and


oxidation resistance, polishability, and high-temperature
properties. It is an excellent carbide former and raises hard-
ening temperatures.
• Cobalt (Co) increases the red hardness/high-temperature
properties in steel.
• Copper (Cu) improves steel’s resistance to corrosion.
• Hydrogen (H) causes embrittlement. It is never added to steel
intentionally.
• Manganese (Mn) is a deoxidizer. It adds to hardenability.
• Molybdenum (Mo) is a carbide former. It increases wear,
hardenability, and red hardness.
• Nitrogen (N) is inherently present in all steels, but in amounts
over 0.004% and in combination with other elements, it will
form hard, abrasion-resistant nitrides.
• Nickel (Ni) provides increased hardenability and toughness.
• Phosphorus (P) is usually a residual, but is added to some
free-machining steels.
• Lead (Pb) is sometimes added to improve machining charac-
teristics.
• Silicon (Si) is a deoxidizer. It adds strength, toughness, and
hardenability.
• Sulfur (S) is sometimes added to improve the machinability
of steel, but it reduces ductility, toughness, and weldability.
• Titanium (Ti) is an excellent carbide former and grain re-
finer.
• Vanadium (V) improves strength and toughness because of
its ability to inhibit grain growth during heat treatment. It
is an excellent carbide former, and is often used to improve
high-temperature properties.
• Tungsten (W) promotes red hardness and a dense, fine grain.
It also promotes wear resistance and high-temperature
strength and is an excellent carbide former.
Carbon is considered the most important alloying element in
steel. The amount of carbon in a steel determines the level of hard-
ness and strength that can be developed by each individual grade
after quenching from the hardening temperature. Higher carbon
steels are prone to be more brittle. The weldability of a steel will

44

Ch04.pmd 44 10/29/02, 11:14 AM


Chapter 4: Metallurgy and Engineering Considerations

decrease with a higher carbon level. Table 4-1 shows terms typi-
cally applied to carbon.

Table 4-1. Typical carbon terms


• Graphitic carbon = free carbon in steel or cast iron
• Combined carbon = carbide carbon + matrix carbon
(that is, all carbon other than free carbon)
• Total carbon = graphitic (or free carbon) carbon + combined carbon

Sulfur is usually an undesirable element in steel because it de-


creases ductility and impact strength and also impairs weldability.
Although sulfur is sometimes added in slight amounts to improve
machinability, it is usually a residual element not deliberately
added. Sulfides act as chip breakers and help to facilitate machin-
ing characteristics. Table 4-2 shows the approximate sulfur limits
for different steel manufacturing processes.

Table 4-2. Approximate sulfur limits


for different steel manufacturing processes
Open-hearth, basic oxygen furnace (BOF) steel = 0.040–0.050%
maximum
Electric furnace steel (vacuum degassed) = 0.025% maximum
Argon-oxygen-decarburization (AOD) quality steel = 0.015%
maximum
Electroslag remelting (ESR), vacuum arc remelting (VAR), vacuum
induction melting (VIM) = 0.005 maximum

CRYSTAL STRUCTURES IN HEAT TREATMENT


The microstructure of tool steel, in the solid condition, is made
up of a number of specific crystal structures, or lattices, with at-
oms arranged in regular geometric patterns. Metals are allotropic
compounds, which means that they can change from one crystal-
line state to another without a phase change. Compare this to
water, which can easily go from a solid to a liquid to a gas phase.
Figure 4-3 describes three of the most common crystal structure
arrangements of tool steels and their unique properties and char-
acteristics.

45

Ch04.pmd 45 10/29/02, 11:14 AM


Tool and Die Making Troubleshooter

Figure 4-3. Crystal structures for tool steels in annealed, austenitic, and
martensitic conditions (Wilson 1975).

In the annealed condition, tool steel’s crystal structures, or space


lattices, are made up of a continuous series of body-centered cu-
bic (BCC) cells. When they are heated to temperatures above ap-
proximately 1,300° F (704° C), their lattices rearrange themselves.
The steel transforms to austenite, which has a face-centered cu-
bic (FCC) crystal structure.
When tool steel is heated to the austenitic temperature range
for a particular chemistry during annealing or hardening, its crys-
tal structure changes to FCC. The steel will become nonmagnetic.
The FCC atoms also become more densely packed, and the steel
shrinks as a result.
Martensite develops when a tool steel is properly quenched to a
low temperature from its austenitic temperature range. Untemp-
ered martensite is unstable. As martensite begins to develop, the
crystal structure of the steel changes from FCC to a less dense
body-centered cubic tetragonal (BCCT) structure. This causes the
steel to “grow.” This growth creates stresses and makes the steel
hard. The change to martensite also causes the steel to regain its
magnetic properties (Wilson 1975; American Society for Metals
1948; Stoughton et al. 1953).

46

Ch04.pmd 46 10/29/02, 11:14 AM


Chapter 4: Metallurgy and Engineering Considerations

EFFECTS OF CARBIDES
Carbides in tool steels are microscopic compounds of carbon
and one or more metallic elements. Carbides are very hard com-
pared to the matrix or cross section of the steel in which they are
embedded. It is their high level of hardness (Table 4-3), in combi-
nation with their homogeneous dispersion throughout the struc-
ture of steel, which greatly adds to abrasion and/or wear resistance.

Table 4-3. Relative hardness of carbides typically found in tool steels


Carbide Element Rockwell C (HRC)
Iron (Fe) 64–68
Chromium (Cr) 75–80
Molybdenum (Mo) 75–80
Vanadium (V) 85–90
Titanium (Ti) 90–100
Tungsten (W) 75–80
Niobium (Nb) 90–100

Carbides that are too large and/or segregated may have an


embrittling effect on steel. Excessive carbide segregation in the
form of heavy banding and/or clusters, such as shown in Figure 4-4,
may contribute to machining problems and brittle failures of tool-
ing in service. These types of carbides and carbide formations will
likely be present in high-carbon, high-chromium tool steels and
in the more highly alloyed types, such as high-speed steel. Car-
bide segregation at the grain boundaries of tooling steels is par-
ticularly objectionable because of its embrittling effect.
Tool-steel manufacturers must have considerable experience to
know what carbide uniformity to expect and what is objection-
able carbide segregation in barstock and plate sections (Roberts
and Cary 1980).

DIRECTIONAL PROPERTIES
All rolled and forged bars exhibit a grain direction that follows
in the direction of working; that is, the direction the steel was

47

Ch04.pmd 47 10/29/02, 11:14 AM


Tool and Die Making Troubleshooter

Figure 4-4. M2 high-speed steel has massive, clustered, banded carbides


(500x).

elongated (referred to as longitudinal) during hot working. Like


wood, the steel is much stronger and more resistant to breakage
when subjected to “against-the-grain forces” (Figure 4-5) than
“with-the-grain forces” (Figure 4-6). Therefore, it is often impor-
tant in the design of tooling components to orient the material’s
grain flow in the best direction to resist major forces on the tool
and avoid breakage parallel to the grain.

WEAR AND FATIGUE


Wear and fatigue are two of the most common causes of prema-
ture tool-and-die failures. Understanding these failure mecha-
nisms is important if production and part-quality goals are to be
met successfully. Wear can be classified into two distinct types,
abrasive wear and adhesive wear.

48

Ch04.pmd 48 10/29/02, 11:14 AM


Chapter 4: Metallurgy and Engineering Considerations

Figure 4-5. Shear blade with grain direction running perpendicular to the
forces applied.

Figure 4-6. Shear blade with grain direction running in the same direction
as the applied forces.

49

Ch04.pmd 49 10/29/02, 11:15 AM


Tool and Die Making Troubleshooter

Abrasive Wear
Abrasive wear is defined as the change in dimension of a part
that is put under conditions of rubbing, grinding, or wearing
away by friction. Note the erosion from the dotted-line surface
in Figure 4-7. Abrasive wear is characterized by the deforma-
tion and erosion (removal) of material from a combination of pres-
sure and grinding action that often develops between a tool and
workpiece. This type of wear usually results in the dulling of cut-
ting edges and working sections. If not properly addressed, the
loss of critical tool dimensions, required tooling tolerances, and
part quality will follow.

Figure 4-7. Typical abrasive wear (LeGrand 1955).

Problems relating to abrasive wear are usually addressed by:


• increasing the hardness of the tooling;
• selecting/using tooling steels with more wear resistance and/
or more wear-resistant carbides; and/or
• changing operating parameters such as rates (speeds, feeds,
cycle times, etc.) or tool clearances.

Adhesive Wear
Adhesive wear is defined as the removal of material from the
surface of a tool or die by a welding action between the workpiece
and the tool. Note the erosion below and weld pickup (galling)
above the dotted surface line in Figure 4-8. The welding action
causes a buildup of material called “pickup.” The weld or pickup

Figure 4-8. Typical adhesive wear (LeGrand 1955).

50

Ch04.pmd 50 10/29/02, 11:15 AM


Chapter 4: Metallurgy and Engineering Considerations

develops from a combination of friction, heat, and pressure, which


develops, for example, between punches and dies, slides, tooling,
and workpieces.
After repeated cycles, the welded material or pickup will have a
tendency to break out and/or shear off, pulling with it material
from the affected part. Eventually, small cavities, voids, and
grooves develop in the working surfaces of affected tools, dies,
machine slides, etc. This leads to dulling, loss of dimension, changes
in tolerances, poor part quality, chipping, and cracking. Problems
relating to seizing, galling, and metal pickup are advanced stages
of adhesive wear.
Adhesive wear can be controlled by reducing the heat in the
metalworking operation by adding lubricants and making changes
in the operation that result in less heat buildup. Surface rough-
ness of the tooling is also important. Finer surface finishes are
less likely to result in adhesive wear. Using dissimilar steels or
identical steels with different hardness levels are other ways to
avoid problems related to adhesive wear.

Fatigue Failures
Fatigue failures occur from repetitive alternating stress load-
ing (tension to compression to tension) of tools, dies, and ma-
chine parts in the presence of stress intensifiers such as sharp
corners. Failures from repetitive stress loading can develop at
stress levels that are very low compared to single-stress static
loads. A simple change of tool-steel grade or hardness level will
not likely resolve problems relating to fatigue failures. Frequently,
fatigue-failure problems can be minimized by the redesign of tool-
ing to eliminate obvious stress raisers (Sandvik Coromant 1996).

TOOL-STEEL TOUGHNESS
One of the most important properties of a tool steel is its ability
to withstand chipping and cracking in the production operation,
where rapidly applied concentrated stresses often develop. The
ability of a piece of steel to resist chipping and cracking depends
on many factors, such as the carbon and alloy content of the steel

51

Ch04.pmd 51 10/29/02, 11:15 AM


Tool and Die Making Troubleshooter

and its hardness. A steel’s ability to withstand stresses from im-


pact is a function of its impact strength or toughness.
The Charpy impact test is one of the most widely accepted tests
used to determine a steel’s relative toughness. The test consists of
putting a machined specimen (per American Society for Testing
and Materials [ASTM] Specification A-370 or ASTM E-23) into a
horizontal fixture. The specimen is then broken by a pendulum
with a sharp hammer head swinging from a fixed height. The en-
ergy absorbed in breaking the specimen is determined by the height
to which the pendulum swings after breaking through the speci-
men. A steel with a very high impact strength will bend before it
breaks. As a result, the hammer will not be able to climb as high
in its arc when compared to a more brittle specimen that will break
easily. Figure 4-9 is a photograph of a Charpy bar that bent before
it broke, and Figure 4-10 shows a Charpy bar that broke with
very little bending (Bethlehem Steel Corp. 1975).

Figure 4-9. Charpy bar that bent before it broke, indicating high impact
strength (Bethlehem Steel Corp. 1975).

Figure 4-10. Charpy bar that broke with little bending, indicating low tough-
ness (Bethlehem Steel Corp. 1975).

52

Ch04.pmd 52 10/29/02, 11:15 AM


Chapter 4: Metallurgy and Engineering Considerations

Table 4-4 compares the Charpy impact strength of several popu-


lar grades of tool steel at various hardness levels.

Table 4-4. Charpy impact strength comparison


Grade Charpy
Charpy,, Charpy
Charpy,, Charpy
Charpy,,
Temper A2 ft-lb
-lbff
ft-lb S7 ft-lb
-lbff
ft-lb D2 ft-lb
-lbff
ft-lb
°F (°C) HRC (N-m or J) HRC (N-m or J) HRC (N-m or J)
400 (204) 60 80 (109) 58 180 60 43.2 (59)
(244)
600 (316) 56 86 (117) 55 228 58 28.8 (39)
(309)
800 (427) 56 83 (113) 53 179 57 41.5 (56)
(243)
900 (482) 56 76 (103) 52 190 58–60 28.1 (38)
(258)

THERMAL CONDUCTIVITY AND FATIGUE


Thermal conductivity is a measure of a material’s ability to
transfer heat. High values of thermal conductivity enable the heat
generated at the working surface of a piece of steel to be dissi-
pated to the body of the part. Low values contribute to the build
up of tool-contact temperatures, which may subsequently mag-
nify cyclic heating and cooling effects that occur in operations,
such as die casting, and result in thermal fatigue.
Figure 4-11 is a simple demonstration of the difference in ther-
mal conductivity. The copper strip, with its higher thermal con-
ductivity (conducts heat at a higher rate) than iron, will ignite a
match sitting on its surface first.
Heat is conducted through metals by two mechanisms. The first
is the transfer of energy by free electrons in the metal. The second,
called “lattice conductivity,” results from the coupling of the vibra-
tion of atoms in the crystals of the metal. When one end of a steel
bar is heated, the atoms at the heated end are excited to a larger
amplitude of vibration than those at the cool end. Thermal mo-
tion is passed from atom to atom, even though each individual

53

Ch04.pmd 53 10/29/02, 11:15 AM


Tool and Die Making Troubleshooter

Figure 4-11. Heat-conduction demonstation in copper and iron (Lincoln


Electric 1973).

atom remains at its original position. Table 4-5 compares the rela-
tive conduction coefficient of various metals based on silver rated
at 100%. Table 4-6 compares the thermal conductivity of various
steels.
Where heat and accompanying thermal cycling from hot to cold
are significant factors, tools and dies will ultimately fail due to

Table 4-5. Conduction coefficient comparison


Conduction Conduction
Metal Coefficient Metal Coefficient

Aluminum (Al) 49.7% Copper (Cu) 92.7%


Gold (Au) 70.9% Iron (Pure) 16.0%
Iron (Steel) 10.9% Iron (Cast) 11.0%
Molybdenum (Mo) 34.9% Nickel (Ni) 14.3%
Platinum (Pt) 16.8% Tin (Sn) 15.6%
Thermal conduction based on silver rated at 100%. (Lincoln Electric 1973)

54

Ch04.pmd 54 10/29/02, 11:15 AM


Chapter 4: Metallurgy and Engineering Considerations

Table 4-6. Thermal conductivity data for various alloy and tool steels
Thermal Conductivity
Temperature, Btu/ft-hr/ft2/°F
Grade °F (°C) (cal-cm/hr-cm2-°F)
4130 Alloy steel* 212 (100) 24.7 (367.6)
572 (300) 21.6 (321.4)
932 (500) 17.9 (266.4)
4140 Alloy steel* 212 (100) 24.7 (367.6)
392 (200) 24.4 (363.1)
752 (400) 21.7 (322.9)
H13 Tool steel 400 (204) 16.5 (245.5)
900 (482) 16.3 (242.6)
S7 Tool steel** 212 (100) 16.5 (245.5)
420 Stainless steel 75 (24) 13.8 (205.4)
212 (100) 14.5 (215.8)
T1 High-speed steel 350 (177) 12.1 (180.1)
800 (427) 14.0 (208.3)
974 (523) 14.5 (215.8)
(*American Society for Metals 1978)
(** Alloy Digest 1976)

the development of a network of fine cracks. The H13 hot-work


tool-steel die-casting die pictured in Figure 4-12 shows severe heat-
check cracking. These cracks form and propagate from thermal
fatigue that develops as alternate expansion and contraction form
residual tensile stresses in the steel. This type of cracking is com-
monly called heat-check cracking, fire cracking, and/or craze crack-
ing.
Conditions that contribute to heat checking include (Bethlehem
Steel Corp. 1981; Roberts and Cary 1980):
• high operating temperatures with long contact times;
• rapid thermal cycling between temperature extremes;
• hardening tools and dies to excessively high hardness levels;
• grain coarsening from overheating during the hardening op-
eration;

55

Ch04.pmd 55 10/29/02, 11:15 AM


Tool and Die Making Troubleshooter

(a)

(b)

Figure 4-12. (a) A severely checked H13 die-casting die and (b) close-up
of heat-check cracks.

• surface carburization or decarburization during heat treat-


ing;
• residual stresses from heat treating and/or grinding, electri-
cal discharge machining (EDM), and welding; and

56

Ch04.pmd 56 10/29/02, 11:15 AM


Chapter 4: Metallurgy and Engineering Considerations

• poor-quality, hot-work tool steels with poorly spheroidized


annealed structures and/or carbides that are networked in
the grain boundaries.
Some tips for minimizing heat-check problems:
• Avoid excessively high hardness levels in tooling.
• Do not exceed recommended hardening temperatures and
soak cycles.
• Heat treat tooling in controlled-atmosphere, salt, or vacuum
furnaces.
• Use multiple tempering cycles after hardening and quenching.
• Use cryogenic tempering, which works to minimize residual
stresses.
• Stress relieve after severe finish grinding, EDM, and/or weld-
ing.
• Use extra clean and uniform electroslag or vacuum-remelted
steel.
• Purchase steel to restricted microstructural standards (for
example, Die Cast Research Foundation Specification NDCA
#207-90, formerly 01-83-02D).

SURFACE SCALE AND DECARBURIZATION


Decarburization describes the loss of surface carbon on a piece
of steel. “Decarb,” as it is often called, develops when steels are
heated to a temperature above approximately 1,250° F (677° C) in
a furnace environment that is oxidizing or lower in carbon poten-
tial than the steel being heat treated requires for a neutral atmo-
sphere.
Mill decarb is the result of the heating of steels to high tem-
perature levels for various steel-mill operations, such as rolling,
forging, annealing, etc. Mill decarb should be completely removed
during the machining of a part and prior to heat treatment. “Heat-
treating decarburization” develops during thermal cycling when
steels are annealed, normalized, and/or hardened in furnaces that
do not have neutral atmospheres. The photo in Figure 4-13 shows
a tool-steel punch that was severely decarburized during harden-
ing. Note the scale flaking off the punch surfaces. Keep in mind

57

Ch04.pmd 57 10/29/02, 11:15 AM


Tool and Die Making Troubleshooter

Figure 4-13. Severely decarburized tool-steel punch.

that the detrimental effects of decarburization go much deeper


than just visual surface scale.
Always protect tool steels during hardening by using a con-
trolled-atmosphere, vacuum, or salt furnace, or a stainless-steel
foil wrap. A surface layer that becomes decarburized during the
hardening operation transforms to austenite on heating, and then
to martensite on quenching at different times than the parent
metal. Therefore, size changes related to transformation take place
at different times in the heat-treat cycle. This causes the develop-
ment of differential stresses that frequently lead to distortion and/
or quench cracking.
The detrimental effects of decarburization include:

58

Ch04.pmd 58 10/29/02, 11:16 AM


Chapter 4: Metallurgy and Engineering Considerations

• the steel has a soft surface layer that also may be rough and
scaled;
• poor wear in heat-treated tooling as a result of low surface
hardness; and
• added distortion and/or cracking of steels in heat treatment.
It is important to note that decarburization penetrates deeper
than the layer of surface scale shown on the punch in Figure 4-13.
The enlarged photomicrograph shown in Figure 4-14 more clearly
depicts the decarburization that developed on an S5 tool-steel
blanking die. The total depth of decarburization exceeded 0.080
in. (2.03 mm), which includes three distinctive layers: (1) surface
scale approximately 0.010–0.015-in. (0.25–0.38-mm) deep, (2) a
free ferrite layer from surface to interior approximately 0.050–
0.060-in. (1.27–1.52-mm) deep, and (3) a decarburized layer ap-
proximately 0.010–0.020-in. (0.25–0.51-mm) deep under the free

Figure 4-14. Photomicrograph (500×) of an S5 tool-steel die badly decar-


burized during hardening.

59

Ch04.pmd 59 10/29/02, 11:16 AM


Tool and Die Making Troubleshooter

ferrite layer where there is a gradual transition to the base-metal


carbon content.

GRAIN COARSENING FROM OVERHEATING


Quenching from excessively high hardening temperatures is one
of the most common causes of premature tool-and-die failure. The
quenching temperature ranges for different tool steels have been
determined by their manufacturers and should be carefully fol-
lowed.
Proper hardening procedures for tool-and-die steels produce a
refining effect on grain size and uniformity. Fractures of properly
hardened tool steels exhibit a uniform, fine-grained, silky struc-
ture corresponding to No. 8, 9, or 10 on the Shepherd Fracture
Grain Size Standards.
When tool steels are heated to excessively high hardening tem-
peratures, grain coarsening is readily visible on a fresh fracture
surface. Figures 4-15 and 4-16 show how grain size has a direct
effect on a steel’s impact strength or toughness (resistance to chip-
ping and cracking). The fracture line in a coarse-grained material
occurs more simply and, thus with less resistance than the frac-
ture line in fine-grained material. On the other hand, smaller fine
grains cause a frequent change in direction of the fracture line,
thus making the material tougher.
Severely overheated steels are also increasingly prone to auste-
nite retention. When the retained austenite transforms in service

Figure 4-15. Coarse-grained steel: Note the simplicity of the fracture line,
a, which has less resistance to crack propagation.

60

Ch04.pmd 60 10/29/02, 11:16 AM


Chapter 4: Metallurgy and Engineering Considerations

Figure 4-16. Fine-grained steel: Note the more complicated fracture line,
b, where the smaller grains cause a frequent change in the direction of
fracture, making the material less brittle (tougher).

or during grinding, residual stresses develop that frequently con-


tribute to early tool-and-die failures. The effect of quenching from
excessively high hardening temperatures is particularly damag-
ing to high-speed steels. Normal quenching temperatures for high-
speed steels are so high that exceeding them will cause these steels
to melt.

INCIPIENT MELTING
Incipient melting is a common problem when hardening high-
speed steels. High-speed steels like M2, M4, and T15 are prone to
incipient melting because their hardening (austenitizing) tempera-
tures are very close to their melting point, approximately 2,300°
F (1,260° C). Incipient melting is defined as the beginning and/or
the initial stage of melting. Because of this, problems may de-
velop when hardening temperatures and/or soaking times are
employed that are higher and longer than recommended.
Problems related to incipient melting are characterized by er-
ratic Rockwell hardness readings, an uneven wave condition (see
Figure 4-17) that develops on the surface of the part being hard-
ened, and/or sticking between the workpiece and stainless-steel foil
wrap sometimes used to protect the steel from decarburization
during hardening. Incipient melting of the microstructure of high-
speed steel causes brittle failures to develop from grain coarsen-
ing, adverse (eutectic carbide) structures, and/or erratic hardness

61

Ch04.pmd 61 10/29/02, 11:16 AM


Tool and Die Making Troubleshooter

Figure 4-17. Wave condition on the surface of an M2 high-speed steel die


plate due to incipient melting.

conditions. High-speed-steel parts damaged by incipient melting


must be scrapped and remade.
Some tips on avoiding problems related to incipient melting:
• Employ a two-stage preheat cycle. For example, M2 may be
preheated at 1,550° F (843° C) until thoroughly soaked, then
heated to 1,850° F (1,010° C) and thoroughly soaked. Heat-
ing to the proper hardening temperature follows, with a soak
for a few minutes up to 15 minutes, depending on part size.
• Check the furnace calibration to ensure that furnace tem-
peratures are not higher than instrument readings show.

62

Ch04.pmd 62 10/29/02, 11:16 AM


Chapter 4: Metallurgy and Engineering Considerations

• Check the temperature consistency and uniformity in the


furnace chamber.
• Employ two thermocouples to monitor furnace and part tem-
peratures.
• Do not overload the furnace. It is better to run two small loads
rather than one load that pushes or exceeds furnace capacity.
• Monitor hardening temperatures and soak times very care-
fully. Avoid oversoaking even by a few minutes.

RETAINED AUSTENITE
Austenite is a metallurgical term defining the crystal structure
to which hardenable steels transform as they are being heated to
their quenching temperature. After transformation to austenite,
the steels must be quenched to transform to martensite (hard steel).
In carbon and low-alloy steels, austenite transforms fairly com-
pletely to martensite or other low-temperature transformation prod-
ucts. However, as the alloy content of the steel increases, increasing
amounts of austenite do not transform to martensite during the
quench, frequently leaving a dangerously high percentage of un-
stable austenite retained in the structure of the as-quenched steel.
This is especially true of high-carbon, high-chromium, air-hard-
ening, and high-speed steels.
If proper measures are not taken to transform the unstable
retained austenite, dimensional shrinkage will occur. Also, if a
tool with retained austenite is placed in service, the austenite will
transform upon application of service stresses. Stress associated
with the newly formed martensite can cause chipping or cracking
of the tool.
Table 4-7 shows test results developed from the testing of 2 × 2
× 2 in. (50.8 × 50.8 × 50.8 mm) M2 high-speed-steel samples for
retained austenite (Teledyne Vasco 1968). They clearly show how
difficult it can be to transform retained austenite to martensite
and the benefit of multiple tempering cycles.
Some tips for controlling problems of austenite retention
(Payson 1962):
• Avoid higher than recommended austenitizing temperatures.
• Do not oversoak at the austenitizing temperature.

63

Ch04.pmd 63 10/29/02, 11:16 AM


Tool and Die Making Troubleshooter

Table 4-7. Effect of tempering M2 high-speed steel


on austenite retention (Teledyne Vasco 1968)
Tempers at 965° F (518° C) % Retained Austenite
As-quenched 24.0
Single, 2-hour temper 16.1
Double temper, 2 + 2 hours 4.7
Triple temper, 2 + 2 + 2 hours 2.7
Quadruple temper, 2 + 2 + 2 + 2 hours 1.7

• Use the highest tempering temperatures possible, consistent


with desired hardness.
• Use double and triple tempering cycles, especially with air-
hardening steels.
• Subzero treatment is another way to guard against austen-
ite retention.

STRESS-RELATED PROBLEMS
One of the keys to the construction of long-running tools and
dies is an understanding of stress development and control. High
residual-stress levels are often the reason why tooling problems
develop related to distortion and premature chipping and cracking.
The problem with stresses is that if they are not minimized,
those from different sources will combine. When stresses from a
number of sources combine and grow to be equal to or exceed the
ultimate strength of the steel, problems related to bending, bow-
ing, twisting, chipping, and cracking will occur.
The primary causes of stress in tooling are:
• mechanical working (for example, operations such as saw cut-
ting, machining, grinding, alternating loads in service, etc.);
• thermal processing (for example, nonuniform heating and
cooling of a steel part, quenching, welding, rapid thermal
cycling, etc.);
• transformation in heat treating (for example, realignment of
the atoms [phase change] in the steel’s microstructure from
hardening); and
• component geometry (for example, irregular design configura-
tions such as sharp internal and external corners, heavy sec-

64

Ch04.pmd 64 10/29/02, 11:16 AM


Chapter 4: Metallurgy and Engineering Considerations

tions adjacent to light sections, thin wall sections, holes, rough


machine marks, etc.).
Some tips for controlling stresses:
• Stress relieve after machining and before hardening.
• Preheat tools and dies before heating to the hardening tem-
perature.
• Carefully anneal hardened tool steels before they are rehard-
ened.
• Do not short cycle the tempering operation after hardening.
• Use multiple tempering cycles where recommended.
• Stress-relieve temper after severe grinding, welding, and EDM.
• Use air-hardening steels whenever design configurations are
complicated and/or massive.

ORANGE PEEL AND PITTING


During the polishing process, exerted pressures have a burnish-
ing rather than a cutting effect that can be the principal cause of
a localized plastic deformation known as orange peel. Orange peel
is characterized by a rippled appearance, which develops when
polishing pressures exceed the yield strength of the steel at its
surface. Once orange peel appears, there is a tendency to apply
more pressure to eliminate the rippled appearance, and doing so
often results in severe pitting of the steel.
Pits are small depressions that may form when small abrasive
particles are torn away from the surface where polishing pres-
sures exceed the tensile strength. The appearance of pits during
polishing is frequently blamed on defects in the steel, but this
often is not the case. A clue to the source of pits may be their
orientation. If the pits are the result of nonmetallic inclusions
present in the steel, they will usually be randomly oriented and
few in number. However, if pitting is the result of overpolishing,
they will tend to be numerous and spread over most of the pol-
ished surface. Evidence of plastic deformation or orange peel also
may be present if the steel has been overpolished.
While the high speeds and pressures common to mechanical
polishing are the main cause of orange peel and pitting, this type
of surface damage may also develop during hand polishing. The

65

Ch04.pmd 65 10/29/02, 11:16 AM


Tool and Die Making Troubleshooter

best way to avoid the possibility of orange peel and pitting is to


keep polishing pressures to a minimum.
The following metallurgical conditions can work alone or in
combination to facilitate the development of orange peel and pit-
ting during polishing:
• Retained austenite on the steel surface is usually the result
of overheating during hardening, carburizing, and/or insuf-
ficient tempering. Because the retained austenite has lower
strength than martensite, it plastically deforms to result in
orange peel and breaks away more easily to cause pitting than
higher-hardness martensitic surfaces.
• Carbides at grain boundaries result from a combination of
overheating and/or inadvertent carburizing in heat treat-
ment: This condition produces nonuniform hardness and rela-
tively low-strength areas. The steel is also susceptible to
pullout of the carbides, resulting in scratches or so-called
“comet tails.”
• Low hardness of the steel is due to incorrect steel composi-
tion or improper heat treatment. Low hardness makes it
easier to exceed the yield strength of the surface during pol-
ishing, resulting in an orange-peel pattern.
• Nonmetallic inclusions in the steel that pull out can result in
pits. The use of cleaner, more homogeneous steels manufac-
tured via double melting (electroslag remelting [ESR] or
vacuum arc remelting [VAR]) may help the polisher avoid
pitting resulting from “pulled-out” inclusions.
• Failure to properly temper steels after hardening may allow
high levels of stress and retained austenite in the steel to
remain. These two factors can work alone or in combination
to contribute to orange peel and pitting.
If orange peel or pitting occurs, it may be possible to repair the
metal surface with the following procedure:
1. Stress relieve the steel by tempering it at a temperature ap-
proximately 50–100° F (28–56° C) lower than the last tem-
pering temperature used.
2. Remove the defective condition by hand polishing with a fine
stone, stress relieving again, and then hand polishing with a
diamond paste using light pressure.

66

Ch04.pmd 66 10/29/02, 11:16 AM


Chapter 4: Metallurgy and Engineering Considerations

3. Minimize the retained austenite in the steel. If orange peel


or pitting is due to excessive retained austenite, the condi-
tion can usually be alleviated by transforming the retained
austenite using additional tempering and/or subzero pro-
cessing. In either case, an additional temper will be neces-
sary to temper the brittle martensite formed from the first
temper.
4. Subzero treat. If a high enough tempering temperature can-
not be used to condition the retained austenite so that it trans-
forms to martensite, subzero processing (–120 to –300° F [–84
to –184° C]) can be used to effect the transformation. Sub-
zero treatments should always be followed by a redundant
tempering operation to temper brittle martensite formed
during the deep freeze.
Note: The type of polish that can be developed on a piece of
steel to a great degree is a function of hardness—the higher the
hardness, the finer the finish obtainable. If an extremely high lus-
ter is required, the hardness should exceed 54 HRC (Young 1967;
Lement Undated; Bethlehem Steel Corp. Undated).

REFERENCES
Alloy Digest. 1976. Data Sheet for Type S7, February. Upper
Montclair, NJ: Engineering Alloys Digest, Inc.
American Society for Metals. 1948. Metals Handbook 1948 Edi-
tion. Cleveland, OH: American Society for Metals, pp. 193–196.
American Society for Metals. 1978. ASM Handbook, Volume 1
Properties and Selection: Irons and Steels, Ninth Edition. Metals
Park, OH: American Society for Metals, p. 148.
Bethlehem Steel Corp. 1972. Tool Steel Topics. Issue 200, Nov./
Dec. Bethlehem, PA: Bethlehem Steel Corp., pp. 1–5.
——. 1975. Tool Steel Topics. Issue 216, July/Aug. Bethlehem, PA:
Bethlehem Steel Corp., pp. 1–3.
——. 1980. Modern Steels and Their Properties. Bethlehem, PA:
Bethlehem Steel Corp., pp. 20–23, 191–198.

67

Ch04.pmd 67 10/29/02, 11:16 AM


Tool and Die Making Troubleshooter

——. 1981. The Tool Steel Troubleshooter. Handbook 2828-C, Janu-


ary. Bethlehem, PA: Bethlehem Steel Corp., pp. 123–124.
——. Undated. “Polishing Plastics Molds.” Tool steel topics spe-
cial report. Folder #2814. Bethlehem, PA: Bethlehem Steel Corp.
Carpenter Technology. 1967. A Simplified Guide for Spark Test-
ing of Tool and Die Steels. Reading, PA: Carpenter Technology, p.
16.
LeGrand, Rupert. 1955. The New American Machinist’s Hand-
book. New York: McGraw-Hill, pp. 40-2 to 40-5.
Lement, Bernard S. Undated. “Plastic Mold Steels.” Booklet pre-
pared in cooperation with the American Iron and Steel Institute
(AISI) Committee of Tool Steel Producers. Phoenix, AZ: Climax
Molybdenum Co., Division of AMAX, pp. 23–24.
Lincoln Electric Company. 1973. Metals and How to Weld Them,
12th Edition. Cleveland, OH: Lincoln Electric Company, pp. 43–
44, 375–386.
Payson, Peter. 1962. The Metallurgy of Tool Steels. New York: John
Wiley & Sons, Inc., pp. 34–68.
Roberts, George A. and Cary, Robert A. 1980. Tool Steels, 4th Edi-
tion. Metals Park, OH: American Society for Metals, pp. 200, 564.
Sandvik Coromant. 1996. Modern Metal Cutting. Sandviken, Swe-
den: Sandvik Coromant, pp. IV-2 through IV-7.
Stoughton B., Butts, A., and Bounds, A. M. 1953. Engineering
Metallurgy. New York: McGraw-Hill, p. 26.
Teledyne Vasco. 1968. High-speed Steels. Latrobe, PA: Teledyne
Vasco, p. 15.
Timken Steel. 1986. Practical Data for Metallurgists. July. Can-
ton, OH: Timken Steel, pp. 90–91.
United States Steel. 1971. The Making, Shaping, and Treating of
Steel, 9th Edition. Pittsburgh, PA: United States Steel, p. 1,128.
Wilson, Robert. 1975. Metallurgy and Heat Treatment of Steels.
New York: McGraw-Hill, pp. 1–2.
Young, William. 1967. Orange Peel and Pitting—Their Causes . . .
Their Cures. Oakdale, PA: Crucible Steel Co.

68

Ch04.pmd 68 10/29/02, 11:16 AM


Chapter 5: Tooling Material Selection

5
Tooling Material Selection

This chapter provides detailed information on steels used for


tooling and plastics molding. The discussion includes a general
description of grades in each category and important related chemi-
cal analyses, properties, and suggested uses, along with notes on
some limitations that should be considered.

IDENTIFYING TOOL-STEEL GRADES


Many manufacturers, distributors, and end-users use color cod-
ing to identify and segregate various tool steels so that they are
not easily mixed (Table 5-1).
American Iron and Steel Institute (AISI) and Society of Auto-
motive Engineers (SAE) designations are designed to make it easy
to identify steels by their principal alloying elements and carbon
levels. The first digit in the number indicates the steel category,
and the second digit indicates the approximate percentage of the
predominant alloying element. The last two digits of the designa-
tion identify the carbon content in hundredths of a percent. For
example: 4150 is the SAE designation for a chromium-molybde-
num alloy steel with a 0.50% carbon level.
Table 5-2 categorizes the AISI and SAE classifications for popu-
lar carbon and alloy steels used today, as well as some that are no
longer commercially available (Bethlehem Steel Corp. 1972).

69

Ch05.pmd 69 10/29/02, 11:17 AM


Tool and Die Making Troubleshooter

Table 5-1. Color coding scheme for various tool steels


AISI Grade Color
A2 Blue
D2 White
H13 Black
H13 SMQ Black and White
M2 Purple
O1 Brown
S7 Red
S7 SMQ Yellow
S5 Pink
W1 Green
4140 PH Orange
1018 CRS Gold
410 SS Grey
420 SS Blue/Red

TOOL-STEEL CLASSIFICATIONS
Tool steels are special types of carbon and highly alloyed steels
capable of being hardened to develop physical and mechanical prop-
erties required for applications such as the stamping, blanking,
bending, forming, drawing, cutting, shearing, trimming, molding,
and extruding of other materials. They are available in the form
of bars, plates, castings, and forgings, which are generally sold in
the annealed condition.
Depending on their alloy content, tool steels may be water, oil,
or air hardening. The least in alloy content are water hardening;
those with more alloy content are oil hardening; and those with
the largest amount of alloy content are air hardening. Upon proper
heat treatment, tool steels are characterized by high hardness and
resistance to abrasion. They are produced under stringent melt-
ing and inspection practices to insure that end users receive the
highest quality material possible for tooling applications.
Tool steels are classified under these general characteristics:
water hardening, oil hardening, air hardening, shock resisting,
hot work, high speed, plastic mold, and special purpose. Within
each group are individual grades with unique chemistries and
properties (Bethlehem Steel Corp. 1979).

70

Ch05.pmd 70 10/29/02, 11:17 AM


Chapter 5: Tooling Material Selection

Table 5-2. AISI and SAE steel classifications


Carbon Steels
10XX (Mn 1.00 maximum)
11XX Carbon only (resulfurized)
12XX Carbon only (resulfurized
and rephosphorized)
15XX Carbon (Mn 1.00–1.65 maximum)
Manganese Steels
13XX Mn 1.75
Nickel Steels
23XX Ni 3.50
25XX Ni 5.00
Nickel- chromium Steels
Nickel-chromium
31XX Ni 1.25; Cr 0.65 or 0.80
32XX Ni 1.75; Cr 1.07
33XX Ni 3.50; Cr 1.50 or 1.57
34XX Ni 3.00; Cr 0.77
Molybdenum Steels
40XX Mo 0.20 or 0.25
44XX Mo 0.40 or 0.52
Chromium-molybdenum Steels
41XX Cr 0.50, 0.80, or 0.95;
Mo 0.12, 0.20, 0.25, or 0.30
Nickel- chromium-molybdenum Steels
Nickel-chromium-molybdenum
43XX Ni 1.82; Cr 0.50 or 0.80; Mo 0.25
47XX Ni 1.05; Cr 0.45; Mo 0.20 or 0.35
81XX Ni 0.30; Cr 0.40; Mo 0.12
86XX Ni 0.55; Cr 0.50; Mo 0.20
88XX Ni 0.55; Cr 0.50; Mo 0.35
93XX Ni 3.25; Cr 1.20; Mo 0.12
94XX Ni 0.45; Cr 0.40; Mo 0.12
97XX Ni 0.55; Cr 0.20; Mo 0.20
98XX Ni 1.00; Cr 0.80; Mo 0.25
Nickel-molybdenum Steels
46XX Ni 0.85 or 1.82; Mo 0.20 or 0.25
48XX Ni 3.50; Mo 0.25

71

Ch05.pmd 71 10/29/02, 11:17 AM


Tool and Die Making Troubleshooter

Table 5-2. (continued)


Chromium Steels
50XX Cr 0.27, 0.40, 0.50, or 0.65
51XX Cr 0.80, 0.87, 0.92, 0.95, 1.00,
or 1.05
Chromium-bearing Steels
50XXX Cr 0.50
51XXX Cr 1.02; C 1.00 minimum
52XXX Cr 1.45
Chromium-vanadium Steels
61XX Cr 0.60, 0.80, or 0.95; V 0.10 or 0.15 minimum
Tungsten- chromium Steels
ungsten-chromium
72XX W 1.75; Cr 0.75
Silicon-manganese Steels
92XX Si 1.40 or 2.00; Mn 0.65, 0.82,
or 0.85; Cr 0 or 0.65
Boron Steels
XX B XX B = boron steels*
Leaded Steels
XX L XX L = leaded steels*
* Xs before the alphabetical character denote principle alloying elements.
(Bethlehem Steel 1972)

Carbon Tool Steel


W1 is water-hardening tool steel supplied with a carbon con-
tent of 0.90/1.00% C. This grade hardens with a hard outside case
and a relatively softer, more ductile inside core.

Cold-work Tool Steel


O1 is a general-purpose, oil-hardening, tool-and-die steel with
good edge-holding ability and high hardness levels.
O6 is a medium-alloy, 1.45% carbon, oil-hardening tool steel. In
the annealed condition, about one third of the carbon is present
in the form of graphitic carbon and the remainder is combined
carbon in the form of carbides. It is the most readily machinable

72

Ch05.pmd 72 10/29/02, 11:17 AM


Chapter 5: Tooling Material Selection

of the oil-hardening tool-steel grades. O6 is good for general-pur-


pose applications requiring resistance to sliding wear and not nec-
essarily maintaining a cutting edge.
A2 is a general-purpose, air-hardening tool steel, which is safe
hardening with low distortion characteristics and high abrasion
resistance. It has a wide range of tooling applications with more
abrasion resistance than the “S” series shock-resistant tool steels
and better toughness and ductility than the “D” series wear
steels.
A6 is a low-alloy, chromium-molybdenum, air-hardening steel
with a balanced combination of toughness, strength, and wear
resistance. It is an excellent choice for intermediate-service cold-
work tools and dies. Because of its low hardening temperature
and air-hardening characteristics, A6 offers safety and dimensional
stability in heat treatment.
D2 is a high-carbon, high-chromium, air-hardening steel for-
mulated to combine excellent abrasion resistance and air-harden-
ing characteristics. It has become the standard against which other
tool steels are measured for abrasion resistance, dimensional sta-
bility in hardening, and air-hardening characteristics.
D3 is a high-carbon, oil-hardening, cold-work tool steel. It pro-
vides maximum wear resistance, high compressive strength, and
deep hardening characteristics. It is used wherever the highest
combination of wear resistance, non-deformation, and hardness
is desired.
D7 is a high-carbon, high-chromium, oil-hardening, cold-work
tool steel with additional carbon and vanadium for maximum abra-
sion resistance. It has good dimensional stability but relatively
low impact strength and toughness. D7 requires slightly higher
than normal hardening temperatures and longer soaking times
during hardening. If properly hardened, D7 exhibits good resis-
tance to softening at elevated temperatures.

Shock-resisting Tool Steel


S1 is a general-purpose, oil-hardening, shock-resisting tool steel.
It has excellent properties for both cold- and hot-work shock ap-
plications. S1 has low carbon content, which gives it good tough-
ness. Chromium and tungsten additions combine to give S1 good

73

Ch05.pmd 73 10/29/02, 11:17 AM


Tool and Die Making Troubleshooter

abrasion resistance and hardenability. S1 is frequently carbur-


ized for additional wear resistance.
S5 is an oil-hardening, shock-resisting steel developed prima-
rily for applications requiring a good combination of ductility and
high hardness levels.
S7 is an air-hardening, shock-resisting tool steel characterized
by excellent impact strength (toughness). S7’s most important
characteristic is its versatility. It is used widely for medium cold-
work tools and dies, plastic molding dies, shear blades, medium
hot-work dies, and the component parts of many products.

Hot-work Tool Steel


H13 is the most widely used air-hardening tool steel for gen-
eral-purpose hot-work applications, die-casting dies, plastic-injec-
tion-mold tooling, and forging tooling. H13 combines good red
hardness, abrasion resistance, and resistance to heat checking.
The optimum working hardness level for H13 is 44–48 HRC for
die-casting tooling, and 40–44 HRC for tooling requiring shock
resistance.

High-speed Tool Steel


M2 is an air-, oil-, or salt-hardening, molybdenum, high-speed
steel with chromium, tungsten, and vanadium. It is a general-pur-
pose, high-speed steel with balanced abrasion and shock resistance
and good red hardness.
M4 is an air-, oil-, or salt-hardening, molybdenum-tungsten,
high-speed steel with high carbon and vanadium content. It has
superior resistance to abrasion when compared to most other high-
speed steels.
M42 is an air- or salt-hardening, molybdenum, high-speed steel
with high carbon and cobalt content. It is characterized by its
ability to be heat treated to 70 HRC. M42 has excellent hot hard-
ness and toughness properties. It also has excellent wear resis-
tance, but because of its chemical composition, it can be ground
with relative ease. M42 should be considered when superior wear
resistance, hot hardness, and toughness characteristics are needed.

74

Ch05.pmd 74 10/29/02, 11:17 AM


Chapter 5: Tooling Material Selection

T15 is an air- or salt-hardening, tungsten-type, high-carbon,


high-vanadium, cobalt, high-speed steel recommended for extreme
abrasion resistance coupled with good red hardness.

Plastic-molding Tool Steel


Type 420 is an oil- or air-hardening stainless steel. It has an
excellent combination of corrosion resistance, abrasion resistance,
and polishability, with a working hardness of 46–50 HRC.
P20 is a prehard (300 HB), medium-alloy mold steel. It is readily
machined in its prehardened condition.

Special-purpose Tool Steel


L6 is a low-alloy, oil-hardening tool steel. It is used for general-
purpose tools and dies where moderate toughness is required at
some sacrifice to abrasion resistance when compared to O1.
Type 4140 prehard is a chromium-molybdenum-alloy steel. It
is delivered fully quenched and tempered to 262–321 HB. High
tensile and yield strengths make this grade very popular for all
kinds of machine parts and tooling applications (Bethlehem Steel
Corp. 1979).

Performance Comparisons
Figure 5-1 graphs the relative abrasion resistance, toughness,
machinability, and grindability of many popular tooling steels at
their normal working hardness levels. Each grade is benchmarked
against a steel grade (left) rated at 100, which is either the best or
close to the best for each particular category. (Note—each grade
has been rated at its normal working hardness [HRC]: M4 62–64,
D2 58–60, M2 60–62, A2 58–60, O1 58–60, O6 58–60, A8 58–60,
S5 56–58, S7 56–58, and H13 42–50.)

TOOL-STEEL SELECTION
The tool-steel selection guide in Table 5-3 can be used to deter-
mine the best grade of tool steel for a given tooling application

75

Ch05.pmd 75 10/29/02, 11:17 AM


Tool and Die Making Troubleshooter

Abrasion resistance

M4 D2 M2
103 100 95
A2
65 O1 A8 O6 S5 S7 H13
50 48 44 44 42 30

Toughness

S7 H13
103 100 S5
90 A8
75
A2
40 D2 M2 M4 O1 O6
20 20 14 14 10

Machinability
O6
125
O1 H13 S7
90 75 75 M2 S5 A2 A8 D2 M4
65 65 65 60 60 60

Grindability

H13
S7 S5
100
92 85 O6 A8 O1
76 75 72
A2
44
D2 M2 M4
10 10 10

Figure 5-1. Tool-and-die steel comparator.

76

Ch05.pmd 76 10/29/02, 11:17 AM


Chapter 5: Tooling Material Selection

Table 5-3. Tool steel selection guide


Application General Best
Anvil S5, S7
Arbors O1, O6, A2, D2
Battering tools (cold) S5 S7
Blacksmith tools S5 S7
Boiler-shop tools S5 S7
Bolt clippers A2 S5
Boring tools M2
Brake dies 4140 HT A2
Broaches—metalworking M2 M4
Burnishing tools A2, D2 M2
Chisels
Blacksmith S5 S7
Chipping S5 S7
Cold working S5 S7
Engraving S5 T1
File cutting D2 M2
Hand S5 S7
Hot working S7 M2
Chuck jaws S5 S7
Clutch parts S5 S7
Collets O1 S5, S7
Concrete breakers S5
Cutters
Form tools M2 M4
Milling M2
Paper D2 M2, M4
Pipe S5, S7 M2
Thread M2 M4
Woodworking O1 M2
Cut-off tools
Cold M2
Hot H13, H21 M2
Dies
Bending O1, O6 A2, D2
Blanking (cold) O1, O6 A2, D2
Blanking (hot) H13, S7 H21
Brake 4140 HT A2
Coining O1 A2, D2
Cold heading A2, D2 M2

77

Ch05.pmd 77 10/29/02, 11:17 AM


Tool and Die Making Troubleshooter

Table 5-3. (continued)


Application General Best
Die casting H13
Die casting (brass) H13 H2
Embossing A2 D2
Extrusion (cold) D2 M2
Extrusion (hot)
Aluminum H13
Copper and brass H21
Forging (hot) H13, S7 H21, H43
Forming (cold) O1 A2, D2
Forming (hot) H13, S7 H21
Gripper (cold) O1, S7 A2, D2
Gripper (hot) S7, H13 H21
Lamination A2 D2, M2
Swaging (cold) S7 D2
Swaging (hot) H13, S7 H21
Thread roll D2 M2
Trimming (cold) O1, A2 D2
Trimming (hot) S7, H13 H21
Wire drawing D2 M4
Drills
Flat, spade M2
Twist M2
Drill bushings S5
Drive rolls D2 D7
Dummy blocks
Hot extrusion H13 H21
End mills M2
Gages A2 D2
Hobs
Cutting M2 T1
Master S7 A2, D2
Knives
Chipper A2, A8 D2, M2
Paper D2 M2
Rotary A8 D2
Shear (cold) A2, D2 S7
Shear (hot) H13, S7 H21
Woodworking A2, D2 M2, 440 Stainless

78

Ch05.pmd 78 10/29/02, 11:17 AM


Chapter 5: Tooling Material Selection

Table 5-3. (continued)


Application General Best
Lathe
Centers A2 D2
Tools M2 M42
Mandrels
Cold working O1, A2 D2
Hot extrusion H13, S7 H21
Molds
Plastic A2, H13, 420, S7
Planer tools M2
Plug gages A2 D2
Pneumatic tools S5 S7
Punches
Center S5 S7
Cold extrusion A2 D2, M2
Cold heading S7, A2, D2, M2
Draw A2, D2, M2
Hot working S7, H13, M2
Piercing A2, S7
Trimming S5, S7 A2, D2
Reamers M2
Rolls
Forming O1, A2 D2, D7
Seaming O1, A2 D2, D7
Screwdriver bits S5
Shaper tools M2
Shear blades
Cold (light gage) A2 D2
Cold (heavy gage) S5 S7
Hot S7, A8
Stamps
Cold S5
Hot S7
Taps O1 M2
(Bethlehem Steel 1978a)

79

Ch05.pmd 79 10/29/02, 11:17 AM


Tool and Die Making Troubleshooter

based on key performance properties such as wear resistance,


impact strength, toughness, red hardness, heat treating, machin-
ability, etc. These recommendations should be used only as a guide.
Individual tool-and-die applications may have unique requirements
and grades other than those listed may need to be used (Bethlehem
Steel Corp. 1978a).

Alloying Elements
Table 5-4 lists the minimum and maximum values for alloying
elements in the more popular tool steels. Unless otherwise indi-
cated, the combined allowable nickel and copper residual is 0.75%
maximum for all the tool steel and high-speed steel grades listed.

Proper ties after Heat Treating


Table 5-5 evaluates key application characteristics for specific
tool-steel grades and offers some heat-treating guidelines to
achieve the best performance.

Hardness Penetration (Hardenability)


Steels with good hardenability will harden more deeply, uni-
formly, safely, and with better dimensional stability and less chance
of cracking than steels with poor hardenability. Those steels that
must be quenched in water to harden are considered to have low
hardenability compared to steels quenched in oil for hardening.
Air-hardening steels are considered to have the best hardenability.
Table 5-6 compares the hardness penetration for various tool steels
(Bethlehem Steel Corp. 1976a; 1977; 1978 b–f, 1981).

Tool Steels for Stamping Dies


The steels listed in Table 5-7 are popular grades for metal-stamp-
ing dies. Each grade’s characteristics and the relative full hard-
ening depth attainable are outlined (Bethlehem Steel Corp. 1976a;
1978b, d; 1981).
The heat treatment of W2 water-hardening material normally
results in a well-defined high-hardness case and a lower-hardness

80

Ch05.pmd 80 10/29/02, 11:17 AM


Ch05.pmd

Table 5-4. Chemical analyses of selected steels


Manganese Phos- Chromium Vanadium Tungsten Molyb-
Carbon % % phorus Sulfur Silicon % % % % denum %
AISI Min./Max. Min./Max. % Max. % Max. Min./Max. Min./Max. Min./Max. Min./Max. Min./Max.
A2 0.95/1.05 0.40/1.00 0.030 0.030 0.10/0.50 4.75/5.50 0.15/0.50 — 0.90/1.40
A7 2.00/2.85 0.20/0.80 0.030 0.030 0.10/0.50 5.00/5.75 3.90/5.15 0.50/1.50 0.90/1.40
A8 0.50/0.60 0.20/0.50 0.030 0.030 0.75/1.10 4.75/5.50 — 1.00/1.50 1.15/1.65
81

D2 1.40/1.60 0.20/0.60 0.030 0.030 0.10/0.60 11.00/13.00 0.50/1.10 — 0.70/1.20


D3 2.00/2.35 0.20/0.60 0.030 0.030 0.10/0.60 11.00/13.50 1.00 1.00 —
D7 2.15/2.50 0.10/0.60 0.030 0.030 0.10/0.60 11.50/13.50 3.80/4.40 — 0.70/1.20
H13 0.32/0.45 0.20/0.50 0.030 0.030 0.80/1.20 4.75/5.50 0.80/1.20 — 1.10/1.75
81

L6* 0.65/0.75 0.25/0.80 0.030 0.030 0.10/0.50 0.60/1.20 — — 0.50


M2 0.78/0.88 0.15/0.40 0.030 0.030 0.20/0.45 3.75/4.50 1.75/2.20 5.50/6.75 4.50/5.50
M4 1.25/1.40 0.15/0.40 0.030 0.030 0.20/0.45 3.75/4.50 3.75/4.50 5.25/6.50 4.25/5.50

Chapter 5: Tooling Material Selection


O1 0.85/1.00 1.00/1.40 0.030 0.030 0.10/0.50 0.40/0.60 0.30 0.40/0.60 —
O6 1.25/1.55 0.30/1.10 0.030 0.030 0.55/1.50 0.30 — — 0.20/0.30
10/29/02, 11:17 AM

S5 0.50/0.65 0.60/1.00 0.030 0.030 1.75/2.25 0.10/0.50 0.15/0.35 — 0.20/1.35


S7 0.45/0.55 0.20/0.90 0.030 0.030 0.20/1.00 3.00/3.50 0.35 — 1.30/1.80
4140(a) 0.38/0.43 0.75/1.00 0.035 0.040 0.15/0.35 0.80/1.10 — — 0.15/0.25
4150(a) 0.48/0.53 0.75/1.00 0.035 0.040 0.15/0.35 0.80/1.10 — — 0.15/0.25
6150(b) 0.48/0.53 0.70/0.90 0.035 0.040 0.15/0.35 0.80/1.10 0.15 — —
*Nickel = 1.25/2.00 (a) 0.35% max. Cu, 0.25% max. Ni (b) 0.35% max. Cu, 0.25% max. Ni, 0.06% max. Mo
(ASTM International 1991, 1992, 1994)
Ch05.pmd

Tool and Die Making Troubleshooter


Table 5-5. Tool steel properties and heat-treating guide
Hardening Tempering
AISI Abrasion Tough- Machin- Temperature Quench Temperature
Grade Chemistry Resistance ness ability
ability** ° F (°C) Type ° F (°C)
W2 C 0.70–1.30; V 0.20 G VG 100 1,450 Water/ 300–600
(788) brine (149–316)
82

O1 C 0.90; Mn 1.20; W 0.50; G G 60 1,475 Oil 350–450


Cr 0.50; V 0.20 (802) (177–232)
O6 C 1.45; Mn 0.80; Si 1.05; G G 125 1,450 Oil 350–450
Mo 0.25 (788) (177–232)
A2 C 1.00; Mn 0.60; Cr 5.25; VG G 60 1,775 Air 400–450*
Mo 1.10; V 0.25 (968) (204–232)
82

A6 C 0.70; Mn 2.00; Si 0.25; G F 65 1,550 Air 400–450*


Cr 1.00; Mo 1.25 (843) (204–232)
D2 C 1.55; Cr 11.50; V 0.90; B F 45 1,850 Air 900–960*
Mo 0.80 (1010) (482–516)
S1 C 0.50; Si 0.75; Cr 1.25; G B 65 1,750 Oil 400–500
W 2.50; V 0.20 (954) (204–260)
10/29/02, 11:17 AM

S5 C 0.60; Mn 0.70; Si 1.85; G B 60 1,600 Oil/water 400–600


Mo 0.45; V 0.20 (871) (204–316)
S7 C 0.50; Mn 0.70; Si 0.25; G B 70 1,725 Air 400–500*
Cr 3.25; Mo 1.40 (941) (204–260)
Ch05.pmd

Table 5-5. (continued)


Hardening Tempering
AISI Abrasion Tough- Machin- Temperature Quench Temperature
Grade Chemistry Resistance ness ability
ability** ° F (°C) Type ° F (°C)
H13 C 0.40; Cr 5.25; Si 1.00; G G 70 1,850 Oil/air 1,050–1,150*
Mo 1.25; V 1.05 (1,010) (566–621)
83

420 C 0.38; Mn 0.40; G G 70 1,800 Oil/air 400–600*


Si 0.40; Cr 13.00 (982) (204–316)
M2 C 0.83; Cr 4.15; VG F 50 2,175 Oil/air 1,000–1,200*
W 6.35; Mo 5.00; V 1.90 (1,191) (538–649)
M4 C 1.27; W 5.50; Cr 4.50; B F 40 2,175 Oil/air 1,000–1,200*
V 4.00; Mo 4.50 (1,191) (538–649)
83

Rating: B-Best, VG-Very Good, G-Good, F-Fair, P-Poor.


* Double tempering recommended

Chapter 5: Tooling Material Selection


**Machinability ratings are related to W-1 as 100%.
(Bethlehem Steel Corp. 1978a)
10/29/02, 11:17 AM
Tool and Die Making Troubleshooter

Table 5-6. Hardness penetration data for various tool steels


W2
Depth of chill, 1/8–3/16 in. (3.2–4.8 mm)
Water quench, chill hardness 68 HRC*
O1
Will through-harden up to 1-3/4 in. (44.5 mm) diameter
Oil quench, 64/65 HRC*
S5
Will through-harden up to 3 in. (76.2 mm) diameter
Oil quench, 62–63 HRC*
A2
Will through-harden up to 4-1/2 in. (114.3 mm) diameter
Air quench, 62–63 HRC*
S7
Will through-harden up to 2-1/2 in. (63.5 mm) diameter
Air quench, 59–60 HRC*
D2
Will through-harden up to 5 in. (127 mm) diameter
Air quench, 61–63 HRC*
M2
Will through-harden up to 1-1/2 in. (38.1 mm) diameter
Air quench, 65–66 HRC*
H13
Will through-harden up to 2-1/2 in. (63.5 mm) diameter
Air quench, 52–54 HRC*
*As-quenched hardnesses
(Bethlehem Steel Corp 1976a, 1977, 1978b–f, 1981)

ductile core. This is not true for the other grades shown in Table
5-8, which are through-hardened. For larger sizes, the S7, A2, and
D2 grades will exhibit a gradual drop in hardness from surface to
center. Actual hardening depths will vary depending on part de-
sign, part mass, hardening temperature, and quenching speed.

Tool Steels for Plastic Molding


Plastic mold steels have to be versatile enough to withstand
varying degrees of alternate heating and cooling from fast-cycling

84

Ch05.pmd 84 10/29/02, 11:17 AM


Chapter 5: Tooling Material Selection

Table 5-7. Popular grades


for automotive metal-stamping dies
Percentage
Element W2 S7 A2 D2
Carbon 1.05 0.50 1.00 1.55
Manganese — 0.70 0.60 —
Molybdenum — 1.40 1.10 0.80
Chromium — 3.25 5.25 11.50
Vanadium 0.20 — 0.25 0.90
Tungsten — — — —
Titanium — — — —
Silicon — 0.25 — —
Total alloy content 1.25 6.10 8.20 14.75

Table 5-8. Characteristics and


relative hardening depth of die steels
W2 Chill depth of approximately 1/8 in. (3.2 mm) per side, 60–62
HRC normal working hardness level, 1,400–1,450° F (760–
788° C) hardening temperature, water quench, wear resistance
good, toughness good.

S7 Fully hardens through a cross-section of 2-1/2 in. (63.5 mm)


round, 56–58 HRC normal working hardness level, 1,725° F
(941° C) hardening temperature, air quench, wear resistance
good, toughness best.

A2 Fully hardens through a cross-section of 4-1/2 in. (114.3 mm)


round, 58–60 HRC normal working hardness level, 1,775° F
(968° C) hardening temperature, air quench, wear resistance
best, toughness good.

D2 Fully hardens through a cross-section of 5 in. (127 mm) round,


58–60 HRC normal working hardness level, 1,850° F (1,010°
C) hardening temperature, air quench, wear resistance best,
toughness fair.
(Bethlehem Steel Corp. 1976a; 1978b, d; 1981)

85

Ch05.pmd 85 10/29/02, 11:17 AM


Tool and Die Making Troubleshooter

molding machines, stress development from impact on parting


surfaces, compressive loads from high clamping pressures, and
wear resistance from metal-to-metal wear and abrasive thermal-
setting resins. Table 5-9 briefly outlines the most popular tooling
steels that have evolved for plastic molding applications. Table 5-
10 presents their characteristics and relative full-hardening depth
(Bethlehem Steel Corp. 1976a, b; 1978f, h; 1981).

Relative Hardness of Tooling Materials


Tool-and-die personnel frequently have to compare the relative
hardness of various tooling materials. Back in the 1930s, 40s, and
50s, a benchmark of comparison was the term “file hard.” A ma-
terial considered file hard was so hard that a hardened and tem-
pered file (usually 60–62 HRC) could not cut into it. In the 1960s
and 70s, many tooling materials were compared to the hardness of
“hard chrome plating” considered to be in the range of 65–75 HRC.
Today, other materials—such as diamond and cubic boron ni-
tride—have supplanted file hard and hard chrome as benchmark
comparisons. Table 5-11 lists the relative hardness of many mate-

Table 5-9. Popular plastics molding steels


Percentage
Element A2 S7 420 SS H13 P-20
Carbon 1.00 0.50 0.38 0.40 0.35
Manganese 0.60 0.70 0.40 — 0.80
Molybdenum 1.10 1.40 — 1.25 0.45
Chromium 5.00 3.25 13.00 5.25 1.70
Vanadium 0.25 — — 1.05 —
Tungsten — — — — —
Titanium — — — — —
Silicon — 0.25 0.40 1.00 0.50
Total alloy 7.95 6.10 14.18 8.95 3.80
(Bethlehem Steel Corp. 1976a, b; 1978f, g; 1981)

86

Ch05.pmd 86 10/29/02, 11:17 AM


Chapter 5: Tooling Material Selection

Table 5-10. Characteristics and


relative hardening depth of plastics molding steels
A2 Fully hardens through a cross-section of 4-1/2 in. (114.3
mm) square, 58–60 HRC normal working hardness level,
1,775° F (968° C) hardening temperature, air quench, wear
resistance very good, toughness good.*
S7 Fully hardens through a cross-section of 2-1/2 in. (63.5 mm)
square, 56–58 HRC normal working hardness level, 1,725° F
(941° C) hardening temperature, air quench, wear resistance
good (but less than A2), toughness very good.*
420 SS Fully hardens through a cross-section of 1-1/2 in. (38.1 mm)
square, 48–50 HRC normal working hardness level, 1,750–
1,850° F (954–1,010° C) hardening temperature, air quench,
wear resistance fair, toughness good, corrosion resistance
very good.*
H13 Fully hardens through a cross-section of 1-1/2 in. (38.1 mm)
square, 46–48 HRC normal working hardness level, 1,850° F
(1,010° C) hardening temperature, air quench, wear resis-
tance fair (but less than S7), toughness very good.*
P-20 Prehardened to a range of approximately 300 Brinell (28–32
HRC). Wear resistance fair (but less than S7), toughness
good.
* Actual hardening depths will vary depending on part design, part mass,
hardening temperature, quenching speed, etc.
(Bethlehem Steel Corp. 1976a, b; 1978f, g; 1981)

rials and surface coatings used for wear enhancement. The hard-
ness information is presented in Knoop (HK) microhardness as
well as the equivalent HRC values.
The HRC conversions are shown because most tooling people
think in terms of that hardness scale. However, the reader is cau-
tioned that the HRC values of 68 or higher are not direct conver-
sions obtained from hardness data tables, but approximations.
Also, comparisons of HRC and Knoop values are nonlinear and,
as a result, the relationship between the two loses accuracy at
very large Knoop numbers. Although these conversion values are
helpful, the only accurate value for hardness is the microhardness
(HK) data.

87

Ch05.pmd 87 10/29/02, 11:17 AM


Tool and Die Making Troubleshooter

Table 5-11. Approximate hardness comparison


of various tooling materials
Approximate Hardness Values
Material/Coating Knoop (HK) Rockwell C (HRC)
Diamond 9,000 99
Cubic boron nitride 5,000 94
Titanium nitride 2,800 87
Vanadium carbide 2,500 85
Silicon carbide 2,400 84
Aluminum oxide 2,100 83
Tungsten carbide 1,900 82
Zirconium oxide 1,500 77
Gas nitrided case 850–1,000 65–70
Chrome plating 900–1,000 68–70
Hardened high speed steels 800–850 63–65
Hardened tool steels 460–690 45–60
Annealed steels 185 20
(Bry Coat, Inc. 2002)

REFERENCES
ASTM International (Formerly American Society for Testing and
Materials). 1991. “Standard Specification for Steel Bars, Alloy,
Standard Grades.” Specification A322. West Conshohocken, PA:
ASTM International.
——. 1992. “Standard Specification for Tool Steel, High Speed.”
Specification A600. West Conshohocken, PA: ASTM International.
——. 1994. “Standard Specification for Tool Steels, Alloy.” Speci-
fication A681. West Conshohocken, PA: ASTM International.
Bethlehem Steel Corp. 1972. “Steel Analyses and Useful Data.”
Booklet 2851. Bethlehem, PA: Bethlehem Steel Corp.
——. 1976a. “Bearcat, AISI S7, Bethlehem, Shock-resisting Tool
Steel.” Folder 2412-B. Bethlehem, PA: Bethlehem Steel Corp.

88

Ch05.pmd 88 10/29/02, 11:17 AM


Chapter 5: Tooling Material Selection

——. 1976b. “AISI Type P20, Bethlehem, Plastic-molding Tool


Steel.” Folder 2797. Bethlehem, PA: Bethlehem Steel Corp.
——. 1977. “BTR, AISI Type O1, Bethlehem, Cold-work Tool
Steel.” Folder 2321-B. Bethlehem, PA: Bethlehem Steel Corp.
——. 1978a. “Tool Steel Selector and Properties Guide.” Booklet
2543-C. Bethlehem, PA: Bethlehem Steel Corp., pp. 2–7.
——. 1978b. “Carbon-vanadium, AISI Type W2, Bethlehem, Wa-
ter-hardening Tool Steel Data Sheet.” Folder 2722-A. Bethlehem,
PA: Bethlehem Steel Corp.
——. 1978c. “Omega, AISI Type S5, Bethlehem, Shock-resisting
Tool Steel.” Folder 2151-E. Bethlehem, PA: Bethlehem Steel Corp.
——. 1978d. “Lehigh, H AISI Type D2, Bethlehem, Cold-work Tool
Steel.” Folder 2323-B. Bethlehem, PA: Bethlehem Steel Corp.
——. 1978e. “AISI Type M2, High-speed Steel.” Folder 2484-B.
Bethlehem, PA: Bethlehem Steel Corp.
——. 1978f. “Cromo-high, V AISI Type H13, Bethlehem, Hot-work
Tool Steel.” Folder 2113-C. Bethlehem, PA: Bethlehem Steel Corp.
——. 1978g. “Bethadur, 420 AISI Type 420, Bethlehem, Plastic-
molding Tool Steel.” Folder 3040-A. Bethlehem, PA: Bethlehem
Steel Corp.
——. 1979. “Tool Steel Topics Mini-course Series.” Booklet 3382,
June. Bethlehem, PA: Bethlehem Steel Corp.
——. 1981. “A-H5, AISI Type A2, Bethlehem, Cold-work Tool
Steel.” Folder 2322-D. Bethlehem, PA: Bethlehem Steel Corp.
Bry Coat, Inc. 2002. “Hardness Conversion Estimate,” cited 2 May.
Available from World Wide Web: <http://www.brycoat.com/
hardness.htm.

89

Ch05.pmd 89 10/29/02, 11:17 AM


Ch05.pmd 90 10/29/02, 11:17 AM
Chapter 6: Tooling Design

6
Tooling Design

Today’s tool design engineers must be “jacks of all trades.” They


must have a good working knowledge of tooling steels, their selec-
tion, basic metallurgy, heat treating, machining, and tool-and-die
failure analysis. Tool design engineers must work to combine tool-
design theory and their computer skills with enough hands-on
experience to fully appreciate the scope and ramifications of tool
construction, application, and maintenance.
This chapter focuses on the fundamentals of proper tool design
and explains how stress raisers frequently cause cracking in heat
treatment and/or premature failure of the tool in service.

MORE THAN JUST MAKING PRINTS


The construction of quality tools and dies begins with tool de-
sign. The advent of computer-aided design (CAD) and computer-
aided manufacturing (CAM) has revolutionized tool design and
construction procedures. Broad experience with all phases of tool-
and-die construction, use, and maintenance make the tool designer
invaluable, not only as a source for accurate drawings, but also
for proactive troubleshooting and problem solving.
Tool designers need to plan ahead and not wait for problems to
develop. They must:
• Minimize design factors that act as stress raisers.
• Give special consideration to tools with massive cross sec-
tions and/or intricate designs, especially in the heat-treat-
ment process.

91

Ch06.pmd 91 10/29/02, 11:17 AM


Tool and Die Making Troubleshooter

• Leave enough time to thoroughly stress relieve, preheat,


harden, quench, and temper.
• Avoid excessively high hardness levels by selecting a range of
hardnesses that best combines requirements for wear resis-
tance and toughness.
• Develop a working relationship with the heat treater.
Tooling cannot be designed on the basis of tool dimensions alone.
The designer must consider the steel to be used as well as its ca-
pabilities and limitations. Factors to consider are whether the tool
must be quenched in water, oil, or air; where tool geometry places
light sections adjacent to heavy sections; and how the tool will be
machined before and after hardening. Dimensional stability also
must be considered to avoid manufacturing and production prob-
lems that can result in expensive repair, production delays, and
downtime.

COMMON DESIGN FAULTS


The importance of good design cannot be overemphasized. Im-
properly designed tools may crack in heat treatment and deliver
zero service life. Or, they may fail prematurely after only a small
portion of anticipated production has been realized. Many tools
that crack during heat treatment should be considered design fail-
ures, and not simply blamed on the heat treater who may have
been asked to harden a poorly designed tool that could not possi-
bly be heat treated successfully.

Sharp Corners
Sharp corners—both internal and external—and sharp-cornered
keyways in tooling act as stress intensifiers. It is always good
machining practice to leave a well-rounded radius in all corners.
Good design practice uses generous filleting to minimize sharp
corners in tooling whenever possible.
Before being released for manufacture, every tool design should
be carefully examined to eliminate sharp corners not essential to
the function of the tool. Sharp corners are the most frequent cause
of premature tool failures and can result in stress-concentration
ratios of 10:1 when compared to the average calculated stress.

92

Ch06.pmd 92 10/29/02, 11:17 AM


Chapter 6: Tooling Design

Light Sections Adjacent to Heavy Sections


During quenching, light sections cool rapidly and harden be-
fore adjacent heavier sections. This sets up differential cooling
and transformation stresses that may exceed the ultimate strength
of the steel. Distortion and/or cracking can result. If such a de-
sign is necessary, heat-treating problems can be minimized by using
air-hardening steels. Keep in mind that problems related to chip-
ping and cracking of tooling in service also can be minimized with
the use of shock-resisting steels.

Problem Holes
Hole placement is an important part of tool design. Blind holes,
threaded holes, or holes improperly placed can be problem areas.
Holes should be designed into tooling, not simply placed. Hole
placement can create areas with thin and weak walls that simi-
larly suffer from differential cooling and transformation stresses
in heat treating and lead to premature cracking. Poor hole place-
ment can lead to areas that are difficult to quench uniformly dur-
ing hardening. This is particularly true if a liquid-quenching steel
is used. If hole designs cannot be improved, use an air-hardening
steel.

Stamp Marks
Excessive stress concentration caused by stamp marks is a major
cause of premature tool failure both in service and in heat treat-
ment. Stamped impressions introduce sharp section changes that
serve to intensify stresses. Avoid sharp, deep-cutting stamps and
bunching stamp impressions together. Whenever possible, use low-
stress stamps with rounded characters. Spread out stamp marks,
and try to minimize the number of letters or numbers. Engraving
using electric pencils, paint, labels, and etching is a better choice.

Surface-finish Defects
Rough machining operations can produce built-in surface
notches and sharp corners that concentrate stress (Bethlehem
Steel Corp. 1981).

93

Ch06.pmd 93 10/29/02, 11:17 AM


Tool and Die Making Troubleshooter

COMMON FAILURE MODES


The following examples are a “rogues gallery” of common tool-
ing failures that occurred when some basic design tenets were
ignored or treated too lightly. The evidence points all too clearly
to tool-design shortcomings, not failures in heat treatment or its
use in production.
For the S7 tool-steel, plastic-injection-mold insert shown in Fig-
ure 6-1, holes, sharp corners and thin wall sections all worked in
combination to weaken it. Note the close-up of the fine, tight crack
radiating from the top of the insert. Sharp corners were natural
stress raisers. The large center hole with only a very thin wall
section adjacent to it was simply too weak to withstand the stresses
of production.
Figures 6-2 and 6-3 show an A2 tool-steel punch and a D2 tool-
steel crimping die, which are representative of many that fail af-
ter very light service. The design of both parts with thin wall
sections, numerous sharp corners, section changes, and cut-out
areas represent dangerous design configurations. Numerous stress
raisers in close proximity severely weaken these tools. Note the
fine, tight cracks radiating out from the sharp internal corners.
Trouble can be expected when hardening a tool with heavy sec-
tions adjacent to light sections. During quenching, light sections
cool rapidly and harden before the adjacent heavy sections, trig-
gering differential cooling stresses that can cause severe distor-
tion and/or cracking of the tool. This is especially true with tool
steels that have to be liquid quenched to develop the required
hardness.
Figure 6-4 shows an O1 tool-steel fixture, six of these either
cracked or severely warped in heat treatment and had to be
scrapped. This tool should have been manufactured from an air-
hardening tool steel with a generous radius used around the base
of the part.
In the case of the A2 tool-steel punch holder in Figure 6-5, sharp
internal corners and undercuts, sharply threaded holes, and thin
wall sections combined to cause cracking during hardening. This
is a dangerous design and would be difficult to manufacture be-
cause there are too many stress raisers concentrated in a very
small area.

94

Ch06.pmd 94 10/29/02, 11:17 AM


Chapter 6: Tooling Design

(a)

(b)

Figure 6-1. (a) S7 tool-steel plastic-injection-mold insert and (b) close-up


view of thin-wall crack .

95

Ch06.pmd 95 10/29/02, 11:18 AM


Tool and Die Making Troubleshooter

Figure 6-2. A2 tool-steel punch with sharp corners and thin wall sections.

Figure 6-3. D2 tool-steel crimping die with sharp corners and thin walls in
close proximity.

96

Ch06.pmd 96 10/29/02, 11:18 AM


Chapter 6: Tooling Design

Figure 6-4. O1 tool-steel fixture with tight crack radiating around the base
of the part.

Figure 6-6 shows one of a series of identical H13 plastic injec-


tion molds that all cracked after very little service. Their failure
was attributed to a combination of heat-treatment irregularities
(insufficient tempering) and service stresses that developed in and
around the top portion of the mold. Numerous stress raisers—
sharp-cornered, blind holes, thin wall sections, and numerous,
significant section changes—likely contributed to these failures.
Figure 6-7 offers two views of an O1 tool steel punch holder. A
fine crack radiates out from a sharp-cornered key slot. The tool
shown was one of 12 identical tools that cracked in heat treatment.
It has numerous stress raisers in close proximity: the sharp-cor-
nered key slot, a sharp undercut, and a very thin wall section.

97

Ch06.pmd 97 10/29/02, 11:18 AM


Tool and Die Making Troubleshooter

Figure 6-5. A2 tool-steel punch holder with cutaway showing stress rais-
ers in the cracked section.

98

Ch06.pmd 98 10/29/02, 11:18 AM


Chapter 6: Tooling Design

Figure 6-6. H13 plastic injection mold and closeup of cracked mold tip.

Figure 6-7. Cracked O1 tool-steel punch holder and closeup of sharp-


cornered key slot and a thin wall section.

99

Ch06.pmd 99 10/29/02, 11:19 AM


Tool and Die Making Troubleshooter

When designs like this cannot be avoided, choose an air-harden-


ing tool steel.
Figure 6-8 is a D2 tool-steel die section that cracked in half
after very light service. Note the presence of numerous stress rais-
ers such as the screw holes, deep cavities, sharp corners, and thin

Figure 6-8. Example of heavy sections adjacent to light, thin wall sections.

100

Ch06.pmd 100 10/29/02, 11:19 AM


Chapter 6: Tooling Design

wall sections. A lower-carbon, shock-resisting tool steel like AISI


S7 would have been a better choice for this tool.
Figure 6-9 is a section from a D2 punch holder that cracked
into three pieces in service. Note the extremely thin wall section
close to the outside diameter of the tool that was not strong enough
to withstand the stresses of production. This tool was eventually

Figure 6-9. End view and cross-section of broken D2 punch holder with
thin wall section.

101

Ch06.pmd 101 10/29/02, 11:19 AM


Tool and Die Making Troubleshooter

redesigned with thicker walls. The material choice was changed


to S7, a low-carbon shock-resisting tool steel with more tough-
ness than D2. Note the cracks radiating from the holes shown in
the end view.
Figure 6-10 shows a large, O1 tool-steel lathe center with poor
hole placement and heavy sections adjacent to light sections. Six
of these parts cracked while they were being hardened and oil

Figure 6-10. O1 tool-steel lathe center that cracked during heat treatment.

102

Ch06.pmd 102 10/29/02, 11:19 AM


Chapter 6: Tooling Design

quenched. Parts like this should be designed with smaller and


less numerous holes and made with an air-hardening tool steel.
Figure 6-11 shows typical cracking through stamp marks. The
S7 tool-steel swaging die in Figure 6-12 is one of a set that cracked
in service. Failure analysis of these parts revealed that deep stamp

Figure 6-11. Two examples of stamp-mark cracking.

103

Ch06.pmd 103 10/29/02, 11:20 AM


Tool and Die Making Troubleshooter

Figure 6-12. Cracked S7 tool-steel swaging die.

marks and less-than-optimum grain orientation of the die steel


likely worked in combination to cause these dies to crack.
The deep stamp markings in Figure 6-13 served to locate the
cracking that occurred. Note how the stress crack passes directly
through the stamped characters. Stamp marks like these tend to
concentrate residual stresses from heat treating and mechanical
stresses from service. Laboratory examination revealed that the
grain flow of the die steel was perpendicular to the direction of
the draw operation.
Like wood, steel is stronger and more resistant to breakage from
against-the-grain forces than with-the-grain forces. Therefore, the
die material will be more resistant to cracking if the grain direc-
tion of the steel runs parallel to the direction of the draw. Re-
placement dies for those in Figure 6-13 used a grain direction
parallel to the direction of the draw and eliminated the deep stamp
marks, which dramatically increased service performance.
Rough machining produces built-in surface notches and sharp
corners that concentrate stresses. All too frequently, rough ma-
chine marks are nucleation sites for crack propagation in service
and heat treatment. Machine operators, die makers, and heat treat-

104

Ch06.pmd 104 10/29/02, 11:20 AM


Chapter 6: Tooling Design

Figure 6-13. Stamp marks plus grain flow caused this die section to fail.

ers must work together, continually reminding one another of the


inherent dangers of this kind of stress raiser.
Figure 6-14 shows a tooling fixture made from D2 tool steel.
Note that the top portion of this tool broke off during heat treat-
ment.

Figure 6-14. D2 tooling fixture with rough machining marks that led to its
cracking.

105

Ch06.pmd 105 10/29/02, 11:20 AM


Tool and Die Making Troubleshooter

Figure 6-15a shows a stress load dispersed over a wide filleted


area, reducing cracking hazards. Figure 6-15b shows how stresses
can be concentrated in a sharp corner location, thus increasing
cracking hazards.

(a) (b)

Figure 6-15. Comparison of stress load for (a) a wide, filleted corner and
(b) a sharp corner.

The estimated impact data in Table 6-1 emphasize that the wider
the radius in a corner section, the stronger and more crack-resis-
tant the tool. Note that the actual impact strength of tooling will
also depend on steel grade, heat treatment, surface finish, etc.

Table 6-1. Approximate strength comparison


of corner sections with different size radii
Radius at Base Projected Impact
of Notch, in. (mm) Strength, ft-lbf (J)
0.002 (0.05) 4 (5.4)
0.010 (0.25) 7 (9.5)
0.020 (0.51) 16 (21.7)
0.040 (1.02) 19 (25.8)
0.080 (2.03) 22 (29.8)
0.125 (3.18) 24+ (32.5+)

106

Ch06.pmd 106 10/29/02, 11:20 AM


Chapter 6: Tooling Design

Table 6-1 is presented only to demonstrate the effect of notch ra-


dius and should not be used for actual strength projections.
One frequent cause of early cracking in corners and sharp radii
is improper seating of the die insert into its holder and/or the im-
proper use of shims. There must be 100% contact between the
bottom surface of the die insert and the holder, or unequal pres-
sures will be exerted. Figures 6-16 and 6-17 illustrate improperly
seated die inserts—the former without a shim and the latter with
an improperly used shim. When only the edges of the die inserts
are touching, an uneven application of force and excessive deflec-
tions contribute to crack development.

Figure 6-16. Improperly seated die insert.

Figure 6-17. Improper use of shims.

107

Ch06.pmd 107 10/29/02, 11:20 AM


Tool and Die Making Troubleshooter

Careful dimensional control in machining of both inserts and


holders avoids the improper seating situations shown in Figures
6-16 and 6-17. When dimensional characteristics do not allow for
100% contact for seating an insert, the best procedure would be
to machine undersize. True up the tool as required to provide for
proper seating. Then add a shim with the necessary thickness to
cover 100% of the seating area. An alternative approach would
involve annealing the tool, welding to produce a buildup, re-heat
treating, and remachining to dimension.

REFERENCE
Bethlehem Steel Corp. 1981. “The Tool Steel Troubleshooter.”
Handbook 2828-C. Bethlehem, PA: Bethlehem Steel Corp., pp.
5–26.

108

Ch06.pmd 108 10/29/02, 11:20 AM


Chapter 7: Tool Machining and Welding

7
Tool Machining and Welding

The shaping and/or repair of tools are accomplished by machin-


ing, grinding, electrical discharge machining (EDM), and weld-
ing. It is essential to know how these processes can affect the tool
steel’s hardness, structure, and metallurgy to avoid problems. This
chapter provides some insight into these areas and includes a de-
scription of the proper methods for shrink-fitting tool components
into retainers.

MACHINABILITY
Machinability—defined as the measure of the ease with which
metals may be cut—depends on the material’s basic characteris-
tics and variables imparted by the machining process. Material
characteristics include hardness, tensile strength, alloy composi-
tion, microstructure, amount of prior cold working, work-hard-
ening characteristics, and the shape, dimension, and rigidity of
the workpiece. Machine variables can include cutting speed, depth
of cut, tool geometry, and the chosen machining process. With so
many variables, any published machinability rating for a specific
material should be considered merely an approximation and used
only for comparative purposes.

Chemical Composition of Material


Because chemical composition affects a material’s structure,
mechanical properties, and heat-treatment response, it has a major

109

Ch07.pmd 109 10/29/02, 11:21 AM


Tool and Die Making Troubleshooter

influence on machinability. Although the effect of individual


elements may be obscured by other factors, some general
comments can be made about the effect of various groups of
elements:
• Carbide-formers such as chromium, tungsten, molybdenum,
and vanadium tend to decrease machinability by increasing
metal hardness.
• Nickel and manganese, which dissolve in ferrite, increase
hardness and toughness, thereby decreasing machinability.
However, these effects can be overcome by annealing treat-
ments.
• Aluminum and silicon form hard abrasive inclusions that
reduce machinability.
• Chemical elements, such as sulfur, lead, phosphorus, sele-
nium, and tellurium, form soft inclusions that improve ma-
chinability.

Prior Processing
Processes such as hot working (forging, rolling), cold working,
and heat treatment affect machinability. In general, large grain
sizes are preferred for most types of machining operations on
metals. Small grain size is preferred if extra-fine surface finishes
must be achieved. The desired grain size can usually be achieved
by controlling the finishing temperature; that is, the tempera-
ture at which hot working is completed.
Large grain sizes may be desirable in carbon steels. However,
large grain size is not desirable for alloy steels because the result
is higher hardenability that produces higher hardness upon cool-
ing from the hot-working temperature. This higher hardness may
mean poorer machinability unless an adequate annealing treat-
ment is employed.
The effect of hot working on machinability is minor when com-
pared to the effects of chemical composition and annealing treat-
ments.
Cold working can improve the machinability of some steels,
particularly those that are soft and gummy, such as the austenitic
stainless-steel types 304, 305, and 316, and ferritic stainless steels

110

Ch07.pmd 110 10/29/02, 11:21 AM


Chapter 7: Tool Machining and Welding

such as types 405 and 430. Cold working is accomplished by cold


drawing bars of these materials, which improves machinability
by causing the steels to be less ductile, at least on the surface. The
lower ductility promotes clean shearing and chip breakage.
On the other hand, cold working will decrease the machinabil-
ity of alloy steels that are susceptible to work hardening. These
steels should not be subjected to heavy roughing cuts because the
result will be work hardening. In addition, the cutting tool should
not be allowed to rub on these steels because it will cause glazing
that will prevent the cutting tool from entering the workpiece.
Heat treatment may markedly affect machinability. For example,
low-carbon steels (under 0.10% carbon) exhibit better machin-
ability after austenitizing and water quenching. The improvement
comes from the lowered ductility of the steel after quenching, re-
sulting in a reduced tendency of the steel to drag during cutting.
Medium-carbon steels (with carbon and alloy contents in the range
of 0.40–0.50%) are usually annealed or normalized to coarsen the
grain size and break up the ferrite matrix (the solid solution that
is the basic structure of the steel) as much as possible. High-car-
bon steels (carbon and alloy contents in the range of 0.90–1.20%
and higher) must be annealed to break up carbide networks and
produce a spheroidized structure.

Physical Characteristics of Material


Grain size, hardness, microstructure, tensile properties, and
thermal conductivity all affect the machinability ratings of steels.
Hardness alone is not a good indicator of machinability because
other key factors—such as composition, structure, and mechani-
cal properties—are not taken into account in a simple hardness
test.
The microstructure of steels has an important influence on
machinability. Low- and medium-carbon steels machine best with
a lamellar pearlitic microstructure, while high-carbon steels ma-
chine best with a spheroidized microstructure.
The tensile properties of steels that affect machining include
elastic modulus, yield strength, ultimate strength, work-harden-
ing exponent, elongation, and reduction of area. Elastic modulus,

111

Ch07.pmd 111 10/29/02, 11:21 AM


Tool and Die Making Troubleshooter

which is independent of the other tensile properties, determines


the rigidity of the workpiece. All carbon and alloy steels have a
high modulus of elasticity and require less support during ma-
chining to prevent deflection away from the cutting tool than low-
modulus materials such as cast irons. The other tensile properties
are interrelated in their influence on metal cutting, and are an
index of machinability only from the standpoint of comparative
values of strength and ductility for different steel compositions.
As a general rule, as the tensile-strength properties increase,
machinability decreases.
High thermal conductivity is beneficial in terms of a steel’s
machinability because heat generated during machining is rap-
idly conducted away from the cutting zone (Allen 1969).

Chips
The machining process is complicated. It involves a combina-
tion of many factors, including the metallurgy of the metal
workpiece, the speed of the machining operation, the tempera-
tures generated during machining, the pressures imparted to the
metal from the cutting tool, and the cutting tool itself. Despite
the many advancements in metal cutting, chip examination and
evaluation is still one of the best ways to monitor the machining
operation. The shape, dimensions, and color of chips throw con-
siderable light on the nature and cutting conditions of the ma-
chining operation that produced them.
Chip formation is affected by several factors:
• workpiece type, strength, hardness, structure, shape, and size;
• feed and cutting depth;
• cutting speed;
• cutting-fluid application; and
• cutting-tool and edge geometry.

Types
Regardless of the type of machining (for example, turning, mill-
ing, drilling, grinding, etc.), there are essentially three kinds of
chips, or chip combinations, generated from metal-cutting opera-

112

Ch07.pmd 112 10/29/02, 11:21 AM


Chapter 7: Tool Machining and Welding

tions: (1) discontinuous or segmented, (2) continuous, and (3) con-


tinuous with built-up edge.
Discontinuous chips. Figure 7-1 shows discontinuous chips,
which consist either of individual segments that may adhere loosely
to each other after the chip has been formed, or as distinct and
unconnected segments produced by actual fracture of the metal
ahead of the cutting edge.

Figure 7-1. Discontinuous chip (Sandvik Coromant 1996).

Discontinuous chips are usually found when machining brittle


materials, or when cutting ductile materials at very low cutting
speeds with large chip thickness and small rake angles. When the
discontinuous chip is associated with brittle materials (that is,
cast irons), it usually results in a fair surface finish, low power
consumption, and reasonably good tool life. With ductile materi-
als (annealed tool steels and stainless steels), the surface finish is
often poor, cutting temperatures rather high, and tool wear ex-
cessive.
Continuous chips. The continuous chip shown in Figure 7-2
is formed by continuous deformation of the metal ahead of the
tool without fracturing, followed by a smooth flow of the chip up
the tool face.
Continuous chips usually result when cutting ductile materials
(annealed tool steels, stainless steels) at cutting speeds above 200
surface ft/min (1 m/sec). They are associated with low friction
between the tool and chip, and large positive rake angles. The

113

Ch07.pmd 113 10/29/02, 11:21 AM


Tool and Die Making Troubleshooter

Figure 7-2. Continuous chip (Sandvik Coromant 1996).

conditions that produce the continuous chip lead to a good sur-


face finish on the workpiece, good tool life, and minimum heat.
Continuous chips with built-up edge. The continuous chip
with a built-up edge shown in Figure 7-3 is similar to a normal
continuous chip, except for an unstable mass of metal that ad-
heres to the face of the tool as the chip shears past it and moves
up the tool face. Fragments of the built-up edge are continuously
forming and then shedding off onto the finished surface as the
tool progresses through the workpiece.
Continuous chips with built-up edges are frequently formed:
(1) when cutting ductile materials at ordinary cutting speeds with
high-speed tools, (2) when cutting ductile materials without fric-
tion-reducing alloy additives, (3) when there is a lack of cutting
fluids, (4) when using rough-ground tool faces, and (5) when pro-
ducing thick chips.

Figure 7-3. Continuous chip with built-up edge (Sandvik Coromant 1996).

114

Ch07.pmd 114 10/29/02, 11:21 AM


Chapter 7: Tool Machining and Welding

The built-up-edge-type chip results in poor surface finish caused


by fragments continuously shedding on the finished surface from
the built-up edge. Tool wear occurs by cratering on the tool face
at the point of contact with the chip, and by abrasion of the tool
flank in contact with escaping severely work-hardened fragments
of the built-up edge.

Color
Because most of the heat generated in the machining operation
is ideally removed by the chip, chip color is a rough indicator of
cutting temperature and can be used as a rough guide to selecting
the optimum cutting-speed range. For example, when machining
with high-speed steel tools, the chips produced should show some
temper coloring: tan indicates 380° F (193° C); straw, approximately
460° F (238° C); or blue, 580° F (304° C). (See Chapter 8, Table 8-
3.) If they do not show color, this means that speed and feed can
probably be greatly increased, thereby improving the machine’s
production rate.
When machining with carbide or ceramic tools, the chips should
always be highly temper-colored (purple or blue). If they are not,
this is an indication of “under machining.” Cemented-carbide tools
can machine at twice the temperature of high-speed steel.
As a general rule, when setting up a machining operation, the
cutting depth should be maximized first and then the feed. Fi-
nally, cutting speed should be established in accordance with rec-
ommended practices for the tool material, relative workpiece
material, and power. Excessive temperatures generated in machin-
ing are the primary cause of unsatisfactory tool life and limita-
tions on high-speed cutting (Sandvik Coromant 1996).

Machining Alloy and Tool Steels


Alloy steels and tool steels are generally sold in the annealed
condition. Annealing of these steels is done for several reasons:
• to remove hot-working strains,
• to make the steel softer for machining,
• to make it more ductile,

115

Ch07.pmd 115 10/29/02, 11:21 AM


Tool and Die Making Troubleshooter

• to refine its grain structure, and


• to produce a definite microstructure in the metal.
Practically all tool steels (O1, A2, D2, S7, H13, etc.) are an-
nealed to a spheroidized structure, because they machine best in
this condition. On the other hand, most alloy and carbon steels
(1060, 1090, 8620, 4140, 4340, 6150, etc.) are annealed to a lamel-
lar pearlitic structure for best machining. Figures 7-4 and 7-5
depict 0.80% carbon steel in both the lamellar pearlitic and
spheroidized annealed conditions.
Due to their generally high carbon content, tool steels machine
best with a spheroidized structure, which allows a cutting tool to
plow through the soft matrix and literally push the hard spheroi-
dal carbides aside (Figure 7-5). Tool steels with a lamellar pearl-
itic structure (Figure 7-4) are very difficult to machine because
the cutting tool is forced to cut through hard, abrasive, elongated
cementite carbide particles.
Carbon and alloy steels are lower in carbon and alloy content
than tool steels. By comparison, these steels have soft and gummy
structures that are difficult to machine. In this case, the resis-
tance to the tool provided by the lamellar pearlitic microstruc-
ture aids the machining process.

Figure 7-4. Lamellar pearlitic microstructure (Allen 1969).

116

Ch07.pmd 116 10/29/02, 11:21 AM


Chapter 7: Tool Machining and Welding

Figure 7-5. Spheroidized condition (Allen 1969).

Machinability Ratings
The machinability ratings in Table 7-1 are based on: (1) rela-
tive machinability of the specific grade as compared to the ma-
chinability of annealed water-hardening tool steel, W1 (rated at
100%) and (2) the experience of machine operators.
In tool steels, 100% machinability is equivalent to about 30%
machinability in constructional steels. For example, the machin-
ability of W2 tool steel is equivalent to approximately 30% of the
machinability of B1112 construction steel rated at 100%. The steels
listed in Table 7-1 are presented in descending order of machin-
ability: that is, from easiest to most difficult to machine. The
Brinell hardness levels shown are the maximum as-supplied an-
nealed hardness levels (or maximum as-supplied prehard levels
for prehardened steels).

Surface-finish Factors
The surface finish of a part after machining is dependent on a
number of factors, such as the shape and smoothness of the cut-
ting tool, machine feeds, speeds, etc. Surface roughness may be
described as irregularities spaced less than 1/32 in. (0.79 mm) apart

117

Ch07.pmd 117 10/29/02, 11:21 AM


Tool and Die Making Troubleshooter

Table 7-1. Relative machinability


of various annealed alloy and tool steels
AISI Machinability* Maximum Brinell
O6 125 229
W1 95 200
W2 95 200
O1 90 221
L6 85 220
A8 85 235
4140 85 197
4150 85 197
420 Stainless 80 230
4340 85 220
S7 75 230
H13 75 235
4140HT 70 321
4150HT 70 321
S5 65 230
P20 65 341
A2 65 235
M2 HSS 65 245
D2 60 235
M4 HSS 55 250
D3 40 255
D4 40 255
D5 40 255
D7 40 255
* Compared to annealed water-hardening tool steel W1 rated at 100%.
(ASM 1980; ASTM International 1994)

and produced by the action of the cutting tool or abrasive. This


determination may be made from a profile curve (Figure 7-6) de-
veloped with a tracer instrument or directly with a profilometer, a
portable instrument that produces an automatic running average
of the height of the surface irregularities.

118

Ch07.pmd 118 10/29/02, 11:21 AM


Chapter 7: Tool Machining and Welding

Figure 7-6. Rough, jagged line depicting a typical profile of a machined


surface (LeGrand 1955).

Surface roughness is measured in microinches or micrometers


(0.000001 in. [25 ␮m]). The most common method used in the
United States today is root-mean-square average (RMS), and is
the method specified by the American National Standards Insti-
tute (ANSI). RMS is the average height of surface peaks above
the nominal. Calculations on data, such as the surface-profile trace
in Figure 7-6, yield this RMS value. Surface roughness also may
be expressed in other ways, such as roughness arithmetic average
(Ra) deviation from the mean, or centerline average (CLA). The
higher the RMS and Ra value, the rougher or coarser the surface.
Some related terminology:
• Waviness is defined as surface irregularities spaced greater
than 1/32 in. (0.79 mm) apart; that is, greater than simple
“roughness.”
• Flaws are irregularities that occur at one place or at rela-
tively infrequent intervals.
• Lay refers to the direction of a predominant surface pattern.
The following RMS references reflect approximate surface-fin-
ish values common in tool and die applications:
• 250 RMS = rough-machined finish, milled flat and square
bars;
• 125 RMS = standard machine finish, Blanchard ground flat
and square bars;
• 65–75 RMS = considered a good machine finish;
• 35 RMS maximum = surface-ground, flat-ground stock;
• 30 RMS = centerless-ground drill rod;
• 16 RMS = typical, good ground finish;
• 8–10 RMS = very good ground finish;
• 4–8 RMS = chrome-plated surface finish; and
• 2–4 RMS = polished and/or lapped surfaces.

119

Ch07.pmd 119 10/29/02, 11:21 AM


Tool and Die Making Troubleshooter

TOOL-WEAR CLASSIFICATIONS
The following classification of tool-wear types form an impor-
tant baseline for assessing machining operations.
All the criteria of machining—economics, accuracy, removal rate,
surface texture, chip and coolant control, etc.—depend to some
degree on the type of tool wear that develops. By inspecting the
magnified cutting edge and acting upon what that tool-wear pat-
tern indicates, the tool designer can control and extend the useful
life of the tool’s cutting edge.

Milling
The following sections discuss the various types of tool wear
that develop during the milling operation.

Flank Wear
The wear shown in Figure 7-7 appears on the flanks of the cut-
ting edge, mainly from abrasion, and is the most normal type.

Figure 7-7. Flank wear (Sandvik Coromant 1996).

Maintaining safe, progressive flank wear is often the ideal. In the


end, excessive flank wear will lead to poor surface texture, inac-
curacy, and increasing friction as the edge changes shape. Some
solutions are to reduce cutting speed and/or select a more wear-
resistant grade.

Notch Wear
Notch wear, shown in Figure 7-8, can occur on either the trailing
edge or the leading edge of the cutting tool. When on the trailing

120

Ch07.pmd 120 10/29/02, 11:21 AM


Chapter 7: Tool Machining and Welding

edge, it is considered typical adhesion wear but also can be due, to


some extent, to the oxidation wear mechanism. The notch will be
formed where the cutting edge and material part. Wear is thus
very localized at the end of the cut, where air (oxidation) can get
to the cutting zone. Notch wear on the leading edge is mechanical
and common with harder materials. Excessive notch wear affects
the surface texture of the workpiece in finishing and eventually
weakens the cutting edge of the insert. Some solutions are to re-
duce the cutting speed and/or select a more wear-resistant grade.

Figure 7-8. Notch wear (Sandvik Coromant 1996).

Crater Wear
Crater wear (Figure 7-9) on the chip face can occur due to abra-
sion and diffusion wear mechanisms. The crater is formed when
the chip face—either by hard-particle-grinding action or through
diffusive action (triggered by high heat and pressure) between
the chip and tool material at the hottest part of the chip face—
wears a crater in the edge of the cutter insert. Hardness, hot hard-
ness, and minimal affinity between materials minimizes the

Figure 7-9. Crater wear (Sandvik Coromant 1996).

121

Ch07.pmd 121 10/29/02, 11:21 AM


Tool and Die Making Troubleshooter

tendency for crater wear. Excessive crater wear changes the ge-
ometry of the cutting edge and can deteriorate chip formation,
change cutting-force directions, and weaken the edge. Some solu-
tions are to: reduce cutting speed and feed; apply coolant prop-
erly; and/or select positive insert geometry.

Plastic Deformation
Plastic deformation (Figure 7-10) is the result of combined high
temperature and high pressure on the cutting edge as high speeds
and feeds and hard workpiece materials create compression and
heat. Edge bulging of the cutting-tool insert is typical and will
lead to even higher temperatures, geometry deformation, chip-
flow changes, etc., until a critical stage is reached. Some solutions
include reducing the cutting speed and feed, and/or selecting a
harder material grade with better resistance to plastic deforma-
tion.

Figure 7-10. Plastic deformation (Sandvik Coromant 1996).

Thermal Cracking
The thermal cracking shown in Figure 7-11 is primarily fatigue
wear due to thermal cycling. Cracks form perpendicular to the
cutting edge and pieces of tool material between the cracks are
pulled out, leading to rapid breakdown and edge failure. Varying

Figure 7-11. Thermal cracking (Sandvik Coromant 1996).

122

Ch07.pmd 122 10/29/02, 11:21 AM


Chapter 7: Tool Machining and Welding

chip thickness also affects temperatures throughout the cut. The


application of cutting fluid can often be detrimental to metal cut-
ting, since the fluid will amplify the temperature variations be-
tween when the cutting tool is in or out of the cut. One solution is
to select a tougher grade material with better resistance to ther-
mal fatigue and shock.

Mechanical-fatigue Cracking
Figure 7-12 depicts mechanical-fatigue cracking, which can take
place when cutting-force shocks are excessive. It is fracture due
to continual variations in load, where the load in itself is not large
enough to cause fracture. The start of the cut and variations in
cutting-force magnitude and direction may be too much for the
strength and toughness of the insert. These cracks are mainly
parallel to the cutting edge. Some solutions are to: reduce the
feed rate; select a tougher grade material; change the cutter posi-
tion; and/or improve stability.

Figure 7-12. Mechanical fatigue cracking (Sandvik Coromant 1996).

Chipping of the Cutting Edge


Chipping of the cutting edge (Figure 7-13) occurs when the edge
line breaks rather than wears. This fatigue-type failure usually
arises from cycles of loading and unloading, which cause particles
of tool material to leave the tool surface. Intermittent cutting is a
frequent cause of this wear type. Careful assessment of the edge
will indicate whether chipping or flank wear is taking place.
Spalling and nicking are variations of this type of edge breakdown.
Some solutions are to: increase the cutting speed; select a tougher
grade material; decrease feed at the beginning of a cut; and/or
improve stability.

123

Ch07.pmd 123 10/29/02, 11:21 AM


Tool and Die Making Troubleshooter

Figure 7-13. Chipping of the cutting edge (Sandvik Coromant 1996).

Insert Breakage or Fracture


Breakage or fracture of the insert (Figure 7-14) can bring a
catastrophic end to the cutting edge. Failure to monitor and/or
correct the failure mechanisms described in this section will al-
ways result in insert breakage or fracture. The change of geom-
etry, weakening of the edge, and rise in temperatures and forces
will eventually lead to major failure of the edge. Some solutions are
to: reduce the feed and depth of cut; select a stronger tool geom-
etry; select a thicker, heavier insert; and/or improve stability.

Figure 7-14. Insert breakage or fracture (Sandvik Coromant 1996).

Built-up-edge Formation
A built-up edge (Figure 7-15) is a phenomenon largely related
to temperature and, therefore, cutting speed. However, it also can
be the result of edge flagging (an inconsistent or intermittent type
of flank wear) and other wear. A built-up edge has a negative ef-
fect on the cutting edge because it changes the geometry of the
tool. Particles of weld material form the built-up edge and break
away from the cutting tool.
The chemical affinity between the tool and the workpiece may
also play an important role (and can cause cratering, as previ-
ously noted). This chemical affinity allows lower temperatures
and high pressures from the machining operation to lead to pres-

124

Ch07.pmd 124 10/29/02, 11:21 AM


Chapter 7: Tool Machining and Welding

sure-welding between the chip and the chip face of the tool. Much
of modern machining takes place at speeds above the built-up edge
range, and many of the newer cutting-tool insert grades are not
as prone to it if used correctly. Surface texture is often the first to
suffer as the built-up edge grows. However, if this type of wear is
allowed to continue, there is a risk of rapid edge breakdown and
fracture. Some solutions are to: increase cutting speed; change to
coated and tougher inserts; and/or apply coolant generously
(Sandvik Coromant 1996).

Figure 7-15. Built-up edge (Sandvik Coromant 1996).

Grinding Hardened Tool Steels


Thermal and transformation stresses that develop from heat
generated during the grinding operation can easily damage the
surface of the part being ground. These stresses can be minimized
by: using grinding fluids that act as both lubricants and coolants
to dissipate heat and avoid stress development; using less rapid
metal-removal rates; and grinding with a wheel of proper grit size
that is not glazed or loaded with grit.
Grinding has often been described as a high-speed milling op-
eration (Figure 7-16) with the grinding-wheel grains compared to
the sharp cutting edges of a cutting tool. These grains have ir-
regular shapes, and their sharp cutting edges and the down-feed
wheel pressures work in combination to generate high tempera-
tures at the point of contact on the surface of the steel. Even though
much of the heat produced is carried away by the metal being
removed and absorbed by the grinding wheel itself, tremendous
heat is still infused into the part being ground.
The heat generated in grinding can result in one or more of the
following, depending on the specific grinding conditions:

125

Ch07.pmd 125 10/29/02, 11:21 AM


Tool and Die Making Troubleshooter

Figure 7-16. The sharp grains in a grinding wheel act as cutting tools as
they slice and rip through a piece of steel.

• The hardness of the surface layer will decrease if the grind-


ing temperatures developed exceed the last tempering tem-
perature of the hardened steel.
• High-temperature infusion into the surface will result in
rehardening of the steel. As a result, the ground surface layer
can contain tempered and untempered martensite as well as
retained austenite. Very high stresses that result from this
are frequently the cause of chipping and cracking.
• The most noticeable effect of abusive grinding is “grinder
scorch,” exhibited by many small cracks that can propagate
if the tool containing them is put into service. This will re-
sult in premature tool failures.

Causes of Grinding Cracks


Problems related to grinding cracks on hardened tool steels are
due to either abusive grinding practice, improper heat treatment,
or a combination of both. It is obvious that considerable energy is
required to grind steels in the hardened condition. The heat from
friction and tearing frequently exceeds the hardening tempera-
ture of the steel. The development of high surface temperatures
during grinding causes two undesirable effects on hardened tool
steels: (1) changes in hardness and/or metallurgical structure, and
(2) the development of high internal stresses from which surface
cracks can form.

126

Ch07.pmd 126 10/29/02, 11:21 AM


Chapter 7: Tool Machining and Welding

Grinding cracks like those shown in Figure 7-17 are initiated


as tiny notches that cannot be seen by the naked eye. Under con-
tinued heat and stress loads, these notches propagate into larger
cracks that become visible and damage the service life of the tool
or die.
Grinding cracks can also result from other poor grinding prac-
tices, including grinding with a dull or a loaded wheel, grinding
with a wheel of grit size too fine for the job, and ineffective use of
coolant.
Heat-treatment-related factors include:
• leaving the tool in the as-quenched condition,
• quenching the tool from an excessively high hardening tem-
perature,
• producing a high-carbon carburized case during the harden-
ing operation, and
• improper tempering procedures, which leave either untemp-
ered martensite or an excessive amount of retained austen-
ite that transforms to as-quenched martensite in grinding.

Salvaging Cracked Tools


Initially, grinding cracks are quite shallow—0.002–0.004 in.
(0.05–0.10 mm)—but if not removed prior to service they will lead
to early failure of a tool via crack propagation.
The D2 tool-steel form die insert in Figure 7-18 was scrapped
and had to be remade because it was quenched from a hardening
temperature that was well above the recommended range for this
grade. It was short cycled during the tempering operation and
abusively surface ground. Because of these heat-treatment irregu-
larities, it is possible that the grinding cracks would have appeared
even with more careful grinding practice.
Provided that further metal removal will not make the tool
unusable, grinding cracks can be removed by following these sal-
vage guidelines:
• Retemper before doing any salvage/repair grinding.
• Remove small amounts of metal on each grinding pass; for
example, 0.00025 in. (0.0064 mm).
• Allow time between each grinding cut to avoid heat buildup.

127

Ch07.pmd 127 10/29/02, 11:21 AM


Tool and Die Making Troubleshooter

Figure 7-17. A2 tool-steel die plate with small, tight grinding cracks.

• Use a soft, coarse-grit wheel (36 grain) that produces mini-


mum heat.
• Use fluorescent powder or dye-penetrant testing to check the
progress of crack removal.

128

Ch07.pmd 128 10/29/02, 11:21 AM


Chapter 7: Tool Machining and Welding

Figure 7-18. D2 tool-steel form die insert with severe grinding cracks.

129

Ch07.pmd 129 10/29/02, 11:22 AM


Tool and Die Making Troubleshooter

A note of caution: If the grinding cracks are the result of grind-


ing an improperly heat-treated tool or die, salvage will be impos-
sible, even with the most careful grinding practices. A simple
retempering operation can often correct the heat-treating defi-
ciency of the affected tool and thus improve the chances of a suc-
cessful salvage grinding operation. If ordinary grinding practices
are used for salvage operations, any cracks present will propagate
and enlarge ahead of the grinding wheel (Bethlehem Steel Corp.
1981; Allen 1969).

EDM EFFECTS ON TOOL STEELS


Electric discharge machining (EDM) is a metal-removal pro-
cess in which a controlled electric current crosses the gap between
a submerged electrode and workpiece. The result is an intense
spark that erodes and/or melts away metal particles. The amount
of melted metal and the depth of the heat-affected zone on a tool’s
surface are directly proportional to the electrical current.
The advantage of EDM is that tooling steels in the hardened
condition (60 HRC and higher) can be easily machined to very
close tolerances and very intricate configurations. However, the
tool-and-die maker must keep in mind that the heat-affected zone
created during the EDM process can adversely affect the perfor-
mance of tools and dies.
The heat-affected zone found on typical EDM surfaces—shown
at 500× in the photomicrograph in Figure 7-19—is made up of
the following three zones:
1. Surface layer: A brittle white layer of as-cast steel, 0.0002–
0.0050 in. (0.005–0.127 mm) deep, will develop. This layer is
characterized by high carbon content, a coarse and granular
structure, 70 HRC or higher in hardness, and microcracks
from 0.0001–0.0003 in. (0.003–0.008 mm) deep.
2. First subsurface: A brittle rehardened, untempered zone,
0.0002–0.0050 in. (0.005–0.127 mm) deep, this as-quenched
martensite will vary in hardness depending on the steel.
Hardness of 65 HRC in steels containing 1% carbon or more
is common.

130

Ch07.pmd 130 10/29/02, 11:22 AM


Chapter 7: Tool Machining and Welding

Figure 7-19. Computer-enhanced photomicrograph of an S7 tool-steel


punch that underwent EDM (500×).

3. Second subsurface: A soft, gradiently tempered zone, 0.0002–


0.0300 in. (0.005–0.762 mm) deep, with hardness of 46–48
HRC, this subsurface will gradually increase in hardness as
unaffected base metal is reached.
Some precautionary measures to consider:
• Use a roughing cut followed by a low-powered finishing cut
to minimize the depth of the white layer.
• Always retemper after EDM.
• Lap grind or polish the EDM surface and then stress relieve
or retemper when finished (Bethlehem Steel Corp. 1979a,
1980).

Wire Burning Hardened Tool Steel


Large-capacity, close-tolerance EDM machines are increasingly
being used to wire-burn tool and die components out of tool-steel
bar and plate sections. It is not uncommon for die makers to cut
mating punches and dies out of one prehardened tool-steel blank.
A number of concerns relating to the premature cracking of parts
either during or subsequent to being wire burned in this way have

131

Ch07.pmd 131 11/13/02, 10:58 AM


Tool and Die Making Troubleshooter

been investigated. These investigations have shown that a very


high percentage of the cracking can be minimized if a more care-
ful and deliberate focus is placed on optimizing heat treatment.
Tool-and-die personnel must keep in mind that it is easier to
control the quality of hardening, quenching, and tempering in a
relatively small rough-machined tool-steel punch, for example,
than it is in a relatively large tool-steel blank from which numer-
ous punches will be wire burned. The large tool-steel plate—3 ×
15 × 22 in. (76 × 381 × 559 mm) long—in Figure 7-20 illustrates
this point. As blanks like this increase in size and mass, concerns
for transforming retained austenite, tempering, and stress reliev-
ing become more critical.
One important factor to be considered is the late transforma-
tion of unstable retained austenite after heat treatment. This
transformation triggers the development of significant critical
stress. This source of stress alone and/or working in combination
with other residual stresses can be a major cause of premature
cracking. The possibility of this kind of cracking will increase as

Figure 7-20. Large S7 tool-steel plate, 3 × 15 × 22 in. (76 × 381 × 559


mm) hardened to 56–58 HRC.

132

Ch07.pmd 132 10/29/02, 11:22 AM


Chapter 7: Tool Machining and Welding

prehardened die sections being wire burned increase in size and


mass.
Figure 7-21 shows a punch that cracked after being wire burned
out of a 3 × 15 × 22-in. (76 × 381 × 559-mm) long, S7 tool-steel
plate hardened to 56–58 HRC. Note that the close-up view of the
crack (Figure 7-22) shows propagation close to a set of bushings
that were inserted with a press fit into holes located close to the
outside diameter of the part. This hole placement in terms of de-
sign creates relatively thin wall sections that are relatively weak
and more susceptible to cracking.
The following tips are offered to minimize these concerns:
• Avoid higher than recommended austenitizing temperatures.
• Do not oversoak at the austenitizing (hardening) tempera-
ture.
• Use the highest tempering temperatures possible, consistent
with the desired hardness.
• Use double and triple tempering cycles, especially with air-
hardened steels.
• A stress relief temper is a must for all hardened tool steels
that have been EDM machined. Select a temperature 25–50°
F (14–28° C) lower than the last tempering temperature used.
• Remove via abrasive lapping the hard, brittle, as-cast EDM
white layer and any small microcracks associated with it.

Figure 7-21. A punch that cracked after being EDM wire burned.

133

Ch07.pmd 133 10/29/02, 11:22 AM


Tool and Die Making Troubleshooter

Figure 7-22. Close-up view of crack in Figure 7-21 that propagated close
to a set of press-fit bushings.

Electrolysis Pitting
Corrosive pitting that develops on the surface of tool steels that
are wire EDM machined is a frequent concern to tool, die, and
mold makers. Although slight corrosive pitting may be overlooked,
larger and more frequent pitting might have to be machined or
polished away. Still more severe pitting may sometimes require
remaking a part.
Surface pitting is a growing concern as increasingly large, mas-
sive, prehardened, and expensive tool-steel workpieces are wire
burned to closer and closer tolerances with increasingly intricate
shapes. Although corrosive pits may sometimes look like metal-
lurgical defects in the steel, close examination frequently reveals
them to be the result of electrolysis etching caused by the EDM
operation itself (Figure 7-23).
During the burning operation, a high-voltage electric current is
conducted through a traveling wire electrode to a grounded work-

134

Ch07.pmd 134 10/29/02, 11:22 AM


Chapter 7: Tool Machining and Welding

Figure 7-23. Pit caused by electrolysis etching during wire EDM. Gap of
0.04 in. (1 mm) is shown at 8×.

piece by a dielectric fluid (usually distilled water). As the current


approaches the workpiece, an electric arc is struck in the form of
a high-voltage spark. This spark produces temperatures in excess
of 4,000° F (2,204° C) that instantly melt and partially vaporize
the workpiece. Electrolysis (triggered by stray current in the di-
electric fluid) releases charged ions and particles from the steel
that has been melted and vaporized during the burning opera-
tion. The ions and particles are magnetically attracted back to
the workpiece where they may corrosively oxidize (rust) and etch
the surface with tiny pits.
Although direct current (DC) power sources are best for high
speed, roughing cuts, they create the most severe corrosive pit-
ting. Alternating current (AC) power sources frequently used for
slower, fine skimming reduce but do not eliminate electrolysis pit-
ting.
One new technology currently under patent by Mitsubishi EDM
is called anti-electrolysis (AE). The AE power source combines
voltage modulation, advanced circuitry, sensors, and software to
interface with the cutting program. AE power supplies used for
roughing and skimming promise full electrolysis protection and
to virtually eliminate rusting and pitting of workpiece surfaces.
The following practices will work to minimize and/or eliminate
electrolysis pitting:
• Vigorously flush and filter out all swarf and contaminants.
• Minimize workpiece submersion in the dielectric fluid.

135

Ch07.pmd 135 10/29/02, 11:22 AM


Tool and Die Making Troubleshooter

• Coat the workpiece with a spray wax that resists the elec-
trolysis etching.
• Introduce food additives (such as savon) to the water to neu-
tralize the positive spark. However, slower cutting rates may
result.
• Employ an AE power supply as described previously (Langen-
hurst 2001).

WELDING TOOL STEELS


Skilled workmanship is the single most important factor when
welding tool steels. Stress relief is the key to die life after welding.
The omission or short cycling of preheat, postheat, stress relief,
and tempering is the main cause of cracking problems, along with
failure to completely and properly remove cracks before welding.
Figures 7-24 and 7-25 shows a heat-affected zone on a piece of
hardened (56–58 HRC) S7 tool steel after welding. Below the weld
deposit is a mixed structure consisting of: (1) steel that has been
rehardened from the heat of welding to 58–60 HRC, and (2) a
zone that has been highly tempered to the point where it is very
soft, 35–40 HRC. These two zones may work in combination to
cause chipping and cracking in service.
Tool steels are very susceptible to cracking when they are be-
ing welded. This is because complicated chemistries allow them
to harden very easily as they cool during, or just after, the weld
operation. Thermal and transformation stresses develop that work
in combination to make crack initiation relatively easy. Compli-

Figure 7-24. Weld-affected area during welding of cast iron.

136

Ch07.pmd 136 10/29/02, 11:22 AM


Chapter 7: Tool Machining and Welding

Figure 7-25. Typical heat-affected zone when welding hardened tool steels.

cated tool-and-die designs with numerous stress raisers—such as


sharp corners, section changes, holes, and massive section sizes—
increase these cracking heat-affected zones.
Comprehensive guidelines for weld repair are difficult to estab-
lish because it is impossible to cover all the different situations
that could occur in the field. Every die-welding job is significantly
affected by such factors as the amount and nature of the welding
required, the design of the die, and the size of the die or compo-
nent being welded.
It does not require a metallurgist to weld tool steels properly,
but tool-and-die welders must have an understanding of the basic
metallurgy of tool steels. Take time to lay out and plan each weld-
ing procedure. Allow enough time for the procedure to be carried
out properly.
Welding tool steels in the hardened condition results in low,
nonuniform hardness and a highly stressed weld zone. The safest
way to weld tool steels is to weld them in the annealed condition.
Rehardening and tempering after welding will restore the steel to
the ideal operating hardness, as well as refine and stress relieve the
weld zone.
Although it is always safest to weld tool steels in the annealed
condition, a high percentage of tool steels are welded in the hard-

137

Ch07.pmd 137 10/29/02, 11:22 AM


Tool and Die Making Troubleshooter

ened condition. When this is done, it is of utmost importance to


use proper preparation, preheat, rod selection, and postheat.

Avoiding Common Pitfalls


The following guidelines and precautions will help the welder
avoid the common pitfalls associated with welding tool steels:
• Double check the grade and hardness of the material. The
more complex the chemistry of the steel, and the higher the
hardness of the tool, the more difficult weld repair becomes.
• Do not short cycle the time needed to anneal, preheat, and
postheat.
• Prepare the surface of the workpiece properly by machining,
scorching (burning off dirt and residue with a torch), or grind-
ing. The area where a crack is removed should be U-shaped,
not V-shaped because sharp angles tend to induce cracking.
Take care to grind/machine below any and all existing cracks;
always provide for a depth of at least 1/8–3/16 in. (3.2–4.8
mm) of weld metal after finishing. Leave a finished weld de-
posit of about 1/16 in. (1.6 mm) for grinding.
• Select the proper weld rod. When welding annealed tool steel,
the chemistry match between the base metal and the weld
metal becomes significant. The weld metal and the base metal
should respond similarly to heat treatment and achieve the
desired hardness. Note: Any highly hardened weld deposit
on the annealed material must be annealed before heat treat-
ing the welded part. If the chemistry of the weld rod cannot
be matched to the base metal, select a weld rod that will du-
plicate the hardness and/or desired properties of the base
metal. When welding hardened tool steel, it is important that
the weld metal achieve the same hardness as the base metal
without going through a heat-treatment cycle.
• Use the smallest-diameter electrode, rod, or wire that will do
the job.
• Make sure the electrodes, rods, or wires are clean and dry
and that the surface being welded is scrupulously clean.
• Never weld a tool-steel tool or die at room temperature. Pre-
heat as recommended, and maintain the preheat tempera-
ture as the interpass temperature.

138

Ch07.pmd 138 10/29/02, 11:22 AM


Chapter 7: Tool Machining and Welding

• Furnace preheat and postheat whenever possible. An oven


or gas-fired heating plate is also acceptable.
• Weld to cause minimum heat. Use minimum recommended
arc voltage and amperage. Reduce amperage slightly for sec-
ondary and finishing passes.
• Position the workpiece so the beads are laid slightly upward.
• Use several small stringer beads rather than heavy deposits.
• Brush slag and dirt from the beads frequently. Peen thor-
oughly to obtain plastic deformation immediately after each
bead is deposited. The metal must be hot—approximately
1,000° F (538° C)—and dark red in color.
• After welding, let the die cool slowly to approximately 160° F
(71° C).
• Retemper previously hardened steels. Reheating dies after
they have been welded removes the welding stresses and tem-
pers the rehardened layer of steel next to the weld.

Welding Hardened Material


Tool steels should never be welded cold, especially hardened
materials. If hardened tool steels are being welded, it is impor-
tant that the preheat temperature be below the original temper-
ing temperature. If the preheat temperature is higher than the
original tempering temperature, the hardness of the tool-steel
workpiece will be correspondingly lower. The preheat tempera-
ture should be maintained as the interpass temperature.
Postheat after welding. Allow welded material to cool to about
160° F (71° C), then temper to relieve welding stresses. Previously
hardened tool steel should be tempered 25–50° F (14–28° C) below
the original tempering temperature so as to retain hardness. A sec-
ond tempering cycle may be effective as an additional precaution
against the premature failure of a previously hardened tool, die,
or mold.
Slow cool after welding in still air. Never force cool with an air
blast or liquid.
Welding a fully annealed die section provides the best weld con-
dition. To avoid decarburizing the workpiece, always anneal in a
controlled neutral atmosphere, a vacuum, or a neutral salt fur-
nace. Welding an annealed die results in a hard weld deposit on

139

Ch07.pmd 139 10/29/02, 11:22 AM


Tool and Die Making Troubleshooter

soft base metal, making a further annealing cycle necessary be-


fore final hardening.

Frequent Causes of Failure


Here are the most frequent causes of weld failure:
• failure to correctly identify the material being welded;
• welding under structural-steel welding conditions (and ig-
noring practically all the extra procedures required for weld-
ing tool steel);
• failure to completely remove cracks before welding;
• failure to follow preheat, postheat, and proper peening pro-
cedures; and/or
• excessively long or wide weld beads.

Welding Tips
Here are some other facts to remember when welding hard-
ened tool steels:
• Although hardened tool steels can be successfully welded, the
welding can be more safely accomplished after annealing of
the steel.
• Match the base material being welded with a weld rod that
will deposit with a comparable hardness level.
• Proper preheating prevents weld cracking. Preheat 25–50° F
(14–28° C) below the tempering temperature.
• Maintain the preheat temperature as the interpass tempera-
ture.
• Reheat the part if the temperature falls significantly below
the preheat temperature.
• Peen immediately after each bead is deposited, when the metal
is a dull red color, approximately 1,000° F (538° C).
• After welding, let the die cool slowly to about 160° F (71° C)
or “hand warm.”
• Temper after welding to relieve welding stresses, and stress
relieve the heat-affected zone and the rehardened layer next
to the weld.

140

Ch07.pmd 140 10/29/02, 11:22 AM


Chapter 7: Tool Machining and Welding

Preheat, Peening, Postheat


Almost every publication on tool steels recommends that tools
in either the annealed or hardened condition be carefully preheated
and postheated before and after welding. Tool steels should never
be welded cold or left untempered after being welded.
Why preheat? Because of their high carbon and alloy contents,
which result in high hardenability, many tool steels are self-hard-
ening during the welding process. As a result, high residual stress
levels may develop, making the tool-steel weldment susceptible to
cracking. Preheating will also help to minimize stress develop-
ment and cracking from thermal shock, reduce the risk of distor-
tion due to excessive heat input, and reduce spalling or cracking
due to underbead embrittlement.
Why postheat? Postheating tempers the steel and controls the
rate of cooling to avoid a hard or brittle structure in the heat-
affected zone. It also stress relieves the weldment.
Why peen? Peening serves to densify the weld, distribute
stresses, and partially stress relieve the weld area.
Here are some other tips:
• Elaborate preheating/postheating equipment is not necessary,
although it is desirable to use a furnace, oven, or heat table
whenever possible.
• Preheating and postheating should be performed in the
temperature range recommended by rod suppliers, and that
temperature should be maintained during welding as the
interpass temperature. Note: If hardened tool steel is being
welded, it is important that the preheat or postheat tempera-
ture be below the original tempering temperature to avoid
temper softening.
• Postheating should be performed immediately after welding.
Allow the welded material to cool to about 150° F (66° C),
then temper to relieve welding stresses. As an additional pre-
caution against the premature failure of a previously hard-
ened tool, die, or mold, a second temper after welding is
suggested.
• To be effective, peening must plastically deform the weld
bead. Never peen when marked resistance to deformation

141

Ch07.pmd 141 10/29/02, 11:23 AM


Tool and Die Making Troubleshooter

is observed; that is, under 1,000° F (538° C). Peen immedi-


ately after each bead is deposited (Bethlehem Steel Corp.
1979b).

SHRINK FITTING
The benefit of shrink-fitting tools, dies, guideposts, or pins made
of carbon, alloy, and tool steels into retainers goes beyond improv-
ing the fixturing techniques. The shrink fit, when properly done,
sets up radial compressive stresses in the tools receiving the fit.
These stresses counteract tensile stresses that frequently lead to
chipping and cracking. Thus, the part receiving the fit will be more
serviceable.

Procedure
Many different methods and techniques can be employed to
accomplish a shrink fit, but for the most part, the procedures
employed are based on the equipment available for performing
the job. Most frequently, the retainer is heated to a temperature
sufficient to cause expansion and allow assembly of the insert.
Caution must be taken not to heat the retainer to a temperature
that exceeds the tempering temperature used in its heat treat-
ment.
Sometimes the procedure is to use heat to expand the retainer
and deep freezing to contract the insert for easier shrink fitting.
In the past, most deep freezing for shrink fitting was performed
with dry ice at approximately –120° F (–84° C).
Today, liquid nitrogen is more frequently used because it has a
lower temperature, approximately –300° F (–184° C). In fact, the
very low temperature range of liquid nitrogen has enabled many
shrink-fitting operations to be performed without heating the re-
tainer. There are some reservations, however, that must be consid-
ered. Unless the cooling is performed slowly (that is, temperature
reduced slowly), the liquid-nitrogen temperature may shock the
insert and cause it to crack. Generally, the amount of contraction
of the insert will vary with the subzero temperature; the lower
the temperature, the greater the shrinkage.

142

Ch07.pmd 142 10/29/02, 11:23 AM


Chapter 7: Tool Machining and Welding

For shrink fitting to be effective, the outside dimension of the


insert is usually 0.003–0.004 in./in. (mm/mm) larger than the in-
side diameter of the retainer. This interference will produce high
joint strength by establishing radial pressure on the mutually con-
tacting surfaces. This radial pressure is proportional to the amount
of interference and the modulus of elasticity of each material. As
a result, tensile stresses are set up in the retainer and compres-
sive stresses in the insert. Careful planning of the die assembly
and careful execution of the shrink fitting will determine the suc-
cess of the operation.

Tips
The shrink fitting of tools into retaining rings is common prac-
tice. Shrink fitting not only facilitates the fixturing of tooling com-
ponents into production machinery, it also improves the service
life of tooling. When properly done, shrink fitting’s radial stresses
counteract tensile stresses that develop in service and are fre-
quently the cause of part cracking. The following are tips and
procedures for proper shrink fitting of tooling.
The first requirement is to design the die and retainer properly:
• The retainer is generally manufactured from an alloy steel
capable of hardening to 32–43 HRC. Shock-resisting tool steels
heat treated to 48–52 HRC are recommended for heavy-duty
applications.
• Adequate interference—usually 0.003–0.004 in./in.(mm/
mm)— should be allowed to ensure satisfactory joint strength.
• An interference of less than 0.001 in./in. (mm/mm) does not
provide adequate radial compressive stress on the insert for
die applications. Use of excessive interference—0.006 in./in.
(mm/mm) or greater—can produce die failures from stress
overload.
• Working stresses (plus shrinkage stresses) must not exceed
the strength of the insert and retainer material.
• Components must be designed for the necessary clearance in
the expanded and/or contracted state to be easily assembled.
• Corners and edges should be chamfered and provided with
generous fillets.

143

Ch07.pmd 143 10/29/02, 11:23 AM


Tool and Die Making Troubleshooter

• A tapered design should not be used on either the insert or


the retainer.
• The outside dimension of the retainer should be at least twice
and preferably three times as large as the inside diameter of
the retainer.
• The inside diameter of the retainer and the outside diameter
of the insert should be ground to a smooth finish (approxi-
mately 5–25 µin. [0.13–0.64 µm]) to provide maximum bond
strength. Grind to within 0.0005 in. (0.013 mm) of the de-
sired dimensions.
Next, complete the job carefully:
• Machine accurately and inspect both the retainer and the
insert with care.
• Clean and dry all parts thoroughly prior to shrink fitting.
• Assemble the heated and cooled parts with a minimum of
delay.
• Allow the assembly to slowly cool (and/or warm) back to room
temperature.
• Do not stress relieve.
A part should never be heated to a temperature higher than
the tempering temperature used in heat treatment because this
will reduce hardness (Bethlehem Steel Corp. 1978, 1981).

REFERENCES
Allen, Dell K. 1969. Metallurgy Theory and Practice. Homewood,
IL: American Technical Publishers, Inc., pp. 192–193, 332–333.
American Society for Metals. 1980. Metals Handbook, Ninth Edi-
tion, Vol. 3, Properties and Selection. Metals Park, OH: American
Society for Metals, p. 443.
ASTM International (formerly American Society for Testing and
Materials). 1994. Standard Specification for Tool Steels, Alloy.
Specification A 681, Table 2. West Conshohocken, PA: ASTM In-
ternational.

144

Ch07.pmd 144 10/29/02, 11:23 AM


Chapter 7: Tool Machining and Welding

Bethlehem Steel Corp. 1978. Bethlehem Tool and Die Steel Manual.
Handbook 2531E. Bethlehem, PA: Bethlehem Steel Corp., pp.
93–94.
——. 1979a. Tool Steel Topics, EDM: Back to Basics. Issue 241.
Bethlehem, PA: Bethlehem Steel Corp.
——. 1979b. Welding Tools, Dies, and Molds. Folder 2599-B.
Bethlehem, PA: Bethlehem Steel Corp.
——. 1980. Tool Steel Topics, EDM: Back to Basics . . . Part 2.
Issue 242. Bethlehem, PA: Bethlehem Steel Corp., pp. 8–9.
——. 1981. The Tool Steel Troubleshooter. Handbook 2828-C.
Bethlehem, PA: Bethlehem Steel Corp., pp. 106–110.
Langenhurst, Greg. 2001. “Anti-electrolysis Developments in Wire
EDM.” Modern Machine Shop On-Line. Cincinnati, OH: Gardner
Publications.
LeGrand, Rupert. 1955. The New American Machinist’s Hand-
book. New York: McGraw-Hill, pp. 40-2 to 40-5.
Sandvik Coromant Company. 1996. Modern Metal Cutting.
Fairlawn, NJ: Sandvik Coromant Company, pp. 1-7 to 1-19, 4-2 to
4-27.

145

Ch07.pmd 145 10/29/02, 11:23 AM


Ch07.pmd 146 10/29/02, 11:23 AM
Chapter 8: Heat Treatment

8
Heat Treatment

Studies of premature tool-and-die failure by a major tool-steel


producer, based on thousands of failure analyses, revealed im-
proper heat treating as the cause for failure 70% of the time.
Thus, proper heat treating is the most significant factor in pro-
ducing tools and dies that attain their expected service life.
Although up-to-date equipment, such as vacuum furnaces, is
recommended for tool-and-die heat treatment, many heat treat-
ers with little more than blacksmith-shop equipment can still pro-
duce excellent results. On the other hand, some heat treaters with
state-of-the-art equipment consistently heat treat tools and dies
that crack and fail prematurely due to poor processing.
This chapter provides an overview of the correct ways to per-
form heat treating of tool-and-die steels and a comprehensive pic-
ture of common problems that can develop and how to avoid them.
The couplets in Figure 8-1 illustrate, with a touch of humor,
the challenges from the heat treater’s perspective.

MORE THAN JUST HARDNESS


Heat treating is the focal point of the tool-manufacturing pro-
cess, yet too much emphasis is placed on heat treating for hard-
ness alone. The hardening operation should develop and fine-tune
working hardness levels with the many different engineering and
physical properties that together guarantee optimum tool perfor-
mance. This requires that tooling-related personnel have a basic

147

Ch08.pmd 147 10/29/02, 11:24 AM


Tool and Die Making Troubleshooter

A Heat Treater’s Lament


You can have your fancy steels,
Fussy treatments, pack-anneals.
All that tungsten, chrome, and moly,
Really gets me down, by golly.
Keeping all their innards straight,
Puts my nerves in an awful state.
When I read up on procedure,
Written by some brainy creature.
I find things like austenite,
Silly words that shed no light.
It’s really quite demoralizing,
To try to fathom normalizing.
Soak and quench, you must preheat,
Transformation incomplete.
Watch your grain and scaly skin,
The atmosphere you heat it in.
Double tempers, strain-relieving,
All add burdens to my grieving.
Each new day is full of terror,
Thinking of the chance for error.
Aiming for those Rockwell highs,
When I’m treating costly dies.
Tools I treat with close attention,
Always warp in each dimension.
My brain’s becoming paralyzed,
Trying to get things carburized.
Salt-bath treatment, heat controls,
Baffle even patient souls.
But when the steel does scale or crack,
The things I’m called . . . alas! Alack!
I don’t mind the gripes profane,
It’s just that I have tried in vain.
To thread my way through complications,
Fancy terms, qualifications . . .
I can’t go on, it’s all too much,
Complainers say I’ve lost my touch.
Treat tool steels? Too hard for me;
Today it takes a Ph.D.

Figure 8-1. Poem (Bethlehem Steel Corp. 1950).

148

Ch08.pmd 148 10/29/02, 11:24 AM


Chapter 8: Heat Treatment

understanding of the fundamentals, metallurgy, common prob-


lems, and techniques for hardening tool steels. They should also
understand the fundamental heat-treating thermal cycles as dia-
grammed in Figure 8-2.

Figure 8-2. Heat-treating thermal cycles for tool steels.

Quality hardening ensures the development of stable, uniform,


and fine-grained microstructures, the minimization of stress lev-
els, and uniform hardness levels.
Here are some basic heat-treating tips:
• Prepare complete hardness specifications for each tool.
• Give special consideration to large batches of tooling with
massive cross sections and/or intricate designs.
• Do not overlook the value of stress-relief cycles, prior to fin-
ish machining, and before hardening.
• Do not short cycle. Plan for enough time to thoroughly stress
relieve, preheat, harden, quench, and temper.
• Check and record the as-quenched hardness to confirm the
success of the hardening operations.
• Avoid hardness levels that are too high. Select a range of
hardnesses that best combines the requirements for abra-
sion resistance and needed toughness.
• Get to know the heat treater and become familiar with his
hardening techniques, capabilities, and limitations.

149

Ch08.pmd 149 10/29/02, 11:24 AM


Tool and Die Making Troubleshooter

LEAVE NOTHING TO CHANCE


Every year, hundreds of thousands of dollars worth of valuable
tooling components are carefully designed and machined, only to
be sent to the heat treatment facility with a piece of scrap paper
with minimal information on how to harden them. As a result,
millions of production dollars are lost when parts fail prematurely
due to less-than-optimum heat treatment. Keep in mind that just
a few hours of heat treatment can make futile hundreds of labor
hours spent on tool design and construction.
Figure 8-3 shows an A2 tool-steel crimping die that cracked
when inserts were press fit into small grooves machined into the
outside diameter of the die. Extreme brittleness resulted from
insufficient tempering triggered by the manufacturer’s request

Figure 8-3. Extreme brittleness from insufficient tempering caused crack-


ing of this A2 tool-steel crimping die.

150

Ch08.pmd 150 10/29/02, 11:24 AM


Chapter 8: Heat Treatment

for a hardness (62–64 HRC) that was inappropriately high for the
design configuration of this tool.
Never leave heat treating to chance. Always plan ahead. Select
the right steel for the job. Take the time to develop and provide
detailed instructions for hardening, quenching, and tempering the
tooling. It is absolutely essential that all important tooling re-
quirements and expectations be fully explained to the heat treater.
This means being insistent about such things as:
• selecting proper temperatures and soaking times for austen-
itizing and tempering;
• documenting the as-quenched hardness of the part(s) after
quenching;
• double and sometimes triple tempering of air-hardening tool
steels;
• selecting the absolute minimum temperature and soak cycles
that may be used during tempering; and
• verifying that the results are in accordance with specifica-
tions.
Most of all, understand that good heat treatment takes time.
Do not push for unrealistic turnaround times that force the heat
treater to shorten critical cycles.

Optimize Communications
Communication is one of the keys to quality tool-and-die con-
struction and maintenance. All parties involved in design, con-
struction, heat treating, and maintenance must work together as
a team to communicate requirements and concerns for optimiz-
ing the production performance of tooling.
Figure 8-4 shows an example of the information that should be
documented whenever tooling is sent to a heat treater and re-
turned from heat-treat processing. The first three steps should be
completed by the submitter and the balance by the heat treater.

Temperature References and Heat Colors


Table 8-1 provides a general reference for many of the tem-
peratures and terms associated with tool-and-die construction,
application, and maintenance.

151

Ch08.pmd 151 10/29/02, 11:24 AM


Tool and Die Making Troubleshooter

Heat TTreating
reating Specification Sheet

1. Grade of steel: _________________________________________________


2. Size, number of pieces: __________________________________________
3. Required hardness: _____________________________________________
4. Furnace type, circle one (atmosphere, salt, vacuum, other)
5. Preheat: temperature, soak time: _________________________________
6. Hardening: temperature, soak time: ______________________________
7. Quench: water, oil, air, other (explain): ____________________________
8. As-quenched hardness: _________________________________________
9. 1st temper, HRC: _______________________________________________
2nd temper, HRC: ______________________________________________
3rd temper, HRC: _______________________________________________
10. Subzero treatment: temperature, HRC: ____________________________
11. Retemper after subzero: temperature, HRC: ________________________

Figure 8-4. A well-documented specification sheet.

Heat and temper colors are frequently used during heat treat-
ing, flame hardening, and welding to visually cross-reference tem-
perature levels in steels to control or monitor processing. Table
8-2 shows the temperature levels for various heat colors. The color
references are obviously subject to individual interpretation as
they are perceived by the eye. Table 8-3 shows the temperature
levels for various temper colors.

FURNACE LOADING
Optimizing furnace loading, part arrangements, and support
during heat treatment is critical whenever parts are to be stress
relieved, normalized, annealed, hardened, quenched, and tempered.
Overloaded furnaces or poorly arranged parts are more difficult to
uniformly heat, soak, and cool. Soak, a term used in the heat treat-
ing of metals, means to hold the item being treated at a particular
temperature for a prolonged, but usually specified, period of time.

152

Ch08.pmd 152 10/29/02, 11:24 AM


Chapter 8: Heat Treatment

Table 8-1. Temperatures for various operations and conditions


Reference °F (°C)
Tungsten arc 11,600 (6,427)
Iron welding arc 10,800 (5,982)
Oxyacetylene flame 6,300 (3,482)
MAPP gas* 5,301 (2,927)
Propane flame 5,300 (2,927)
Oxyhydrogen 5,072 (2,800)
Laboratory burner 3,360 (1,849)
Iron melts (liquid) 2,800 (1,538)
Welding and steel melting 2,400 (1,316)
High-speed hardening 2,100–2,400 (1,149–1,316)
Alloy, tool-steel hardening 1,500–1,900 (816–1,038)
Carburizing range 1,700–1,800 (927–982)
Flame-hardening range 1,550–1,600 (843–871)
(4140/4150)
Annealing range 1,400–1,700 (760–927)
Carbon steel hardening 1,350–1,550 (732–843)
Stress relieving 400–1,200 (204–649)
Nitriding range 900–1,000 (482–538)
Blue brittle range 600–800 (316–427)
Water boils 212 (100)
Ice melts 32 (0)
Zero Fahrenheit 0 (–18)
Dry ice –120 (–84)
Subzero treatment –320 to 0 (–196 to –18)
Absolute zero –460 (–273)
(all molecular activity ceases)
* Liquefied petroleum and methylacetylene-propadiene
(Bethlehem Steel Corp. 1936)

The soak may be employed at the hardening temperature to develop


proper austenization or during tempering to make tempering ef-
fective. Improperly supported parts frequently bend under their

153

Ch08.pmd 153 10/29/02, 11:24 AM


Tool and Die Making Troubleshooter

Table 8-2. Temperature levels for various heat colors


Heat Color °F (°C)
Yellow/white 2,500 (1,371)
Yellow 2,300 (1,260)
Bright orange 2,100 (1,149)
Orange 1,600 (871)
Bright red 1,500 (816)
Cherry red 1,400 (760)
Dark (black) red 1,250 (677)
(Bethlehem Steel Corp. 1936)

Table 8-3. Temperature levels for various temper colors


Temper Color °F (°C)
Light gray 700 (371)
Gray 660 (349)
Blue 580 (304)
Blue, black (gunmetal) 500 (260)
Straw, brown 475 (246)
Straw 460 (238)
Tan 380 (193)
(Bethlehem Steel Corp. 1936)

own weight at elevated temperatures. This is particularly impor-


tant when sections being processed have complicated configura-
tions and/or are to be processed in batches. The illustrations in
Figure 8-5 and the following tips can help optimize heat-treat pro-
cessing:
• Never overload the furnace. In an overloaded furnace, the
parts will not be exposed to the proper circulation of the fur-
nace atmosphere. It is also more difficult to bring all the parts
to the required temperature uniformly. When significant lag
times develop between the overall furnace and individual
material temperatures, chances of heat-treat error increase.

154

Ch08.pmd 154 10/29/02, 11:24 AM


Chapter 8: Heat Treatment

Figure 8-5. Good and poor furnace loading techniques.

• Keep parts being processed as far apart as possible to pro-


mote uniformity in heating and cooling. It is always good
practice to use several thermocouples in a furnace to moni-
tor furnace temperature levels for uniformity as well as the
temperatures of the parts themselves.
• Always place all material to be processed on supportive racks
(as in Figure 8-5), again, to achieve a higher degree of unifor-
mity in heating and cooling. Racks are generally fabricated

155

Ch08.pmd 155 10/29/02, 11:24 AM


Tool and Die Making Troubleshooter

from heat-resisting materials, including 300-series stainless


steels. The specific alloys used will depend on anticipated long-
term furnace service conditions. Round bars 1/2–3/4 in. (12.7–
19.1 mm) in diameter provide minimal surface contact with
the heat-treat materials.
• Try to load furnaces with parts of uniform size, dimension,
or mass.

HARDENING AND RELATED STRESSES


It is important to understand the stresses that occur during
heat treatment. Figure 8-6 illustrates a typical hardening proce-
dure for air-hardening D2 tool steel. The hardening cycle begins
with a preheat at 1,200° F (649° C), followed by austenitizing at
1,850° F (1,010° C). After austenitizing, the steel is air quenched
to 150° F (66° C). Quenching is followed by multiple tempers and
possible deep freezing.
When the steel is being heated to the preheat and austenitizing
temperatures, it will expand and begin to develop thermal stresses.
As the furnace drives heat into the steel part, its surface will ex-
pand before the interior, which lags behind in temperature until
the part becomes fully equalized. This differential expansion cre-
ates additional thermal stresses.

Figure 8-6. Hardening cycle for D2 tool steel.

156

Ch08.pmd 156 10/29/02, 11:24 AM


Chapter 8: Heat Treatment

Transformation stresses develop when the Ac1 temperature is


reached. At this point, the microstructure of the steel begins to
contract as austenite is formed. This contraction continues until
the part becomes fully austenitic (that is, when Ac3 is reached
throughout the part). Contraction occurs because austenite is
smaller in volume per unit of mass than an annealed structure.
This contraction, combined with the differential temperature be-
tween the surface and the interior of the part during heating,
results in additional stresses. After Ac3 is reached, the steel con-
tinues to expand on heating to the austenitizing temperature.
The faster the heating rate, the greater the development of
stresses. Note that because of temperature gradients, these
stresses will be greater in parts with complicated design configu-
rations and/or where thin sections are located adjacent to thick
sections. As the part reaches the austenitizing temperature
throughout, and is sufficiently soaked at that temperature, the
stress level goes to zero.
When the part is quenched from the austenitizing temperature,
the cooling of the interior lags behind the surface temperature
and, again, there is a development of thermal stresses as the sur-
face contracts before the interior. The faster the quench rate, the
greater the temperature differential between the surface and in-
terior. As a result, the degree of contraction that occurs at the
surface, as compared to the interior, will be greater and the re-
sidual stresses will increase in magnitude.
When the temperature of the steel during quenching reaches
the point where martensite begins to transform, contraction of
the part is halted. The steel will then begin to grow. Transforma-
tion stresses will develop from this growth and continue to grow
until martensitic formation ends throughout the part. At this
point, the workpiece may begin to contract again with further
cooling.
At the end of the quench, the part will have grown when com-
pared to its size in the annealed condition. This is because mar-
tensite is larger in volume per unit of mass than the annealed
material and the austenite from which it was formed. Differential
stresses may also develop because this growth may begin at dif-
ferent time intervals in the steel, depending on the uniformity of
the quench, the thickness of the material, its design, etc.

157

Ch08.pmd 157 10/29/02, 11:24 AM


Tool and Die Making Troubleshooter

Multiple tempering and subzero quenching operations are im-


portant final steps in the hardening process. They are used to
stress relieve the steel after it is quenched and make the transfor-
mation of austenite to martensite more complete. The result will
be a part that is less susceptible to size change and/or chipping
and cracking during finish grinding, electrical discharge machin-
ing (EDM), welding, or the production operation.

STAINLESS FOIL WRAPS


When atmosphere-controlled or vacuum furnaces are not avail-
able, stainless foil (steel) wrap is frequently used for the surface
protection of tool-steel parts during the hardening operation.
Stainless foil prevents problems related to decarburization of tool
steels for hardening temperatures up to 2,240° F (1,227° C). Stain-
less foil wrap is recommended for air-hardening steels when their
austenitizing temperatures are below 2,000° F (1,093° C), and when
soaking times do not exceed 6 hours. High-temperature foil wrap
is recommended for air-hardening steels that have austenitizing
temperatures up to 2,240° F (1,227° C). Note that foil wrap is rec-
ommended for air-hardening grades only and is applied prior to
the start of heat treating. Oil-hardening tool steels are not nor-
mally wrapped because the foil would have to be removed from
these steels to get a good quench.
Some tips on using foil wrap:
• Workpieces must be wrapped and sealed in the foil with care.
• Do not rip, puncture, or tear the foil wrap.
• Cover sharp corners of the part being treated with small pieces
of foil to keep the sharp corners from puncturing the package.
• Triple-fold all open edges to develop an airtight compartment.
• Heat treat and air quench normally. (Slightly longer-than-nor-
mal soak cycles may be necessary to develop high as-quenched
hardness due to the insulation effect of the foil wrap.)
• It is not necessary to remove the foil wrap from the steels
before quenching. There is usually sufficient heat transfer
through the foil for the quenching of air-hardening steels.
The result will be a hardened part that is free of oxide, scale,
and decarburization.

158

Ch08.pmd 158 10/29/02, 11:24 AM


Chapter 8: Heat Treatment

At times, it may be beneficial to quench large tools to 1,000° F


(538° C) in the foil. After quenching to 1,000° F (538° C), open the
package, remove the part, and air quench to 150° F (66° C). Tem-
per immediately (Bethlehem Steel Corp. 1967; N. E. Slavin and
Co. 1972; Teledyne Sterling Hunt Undated).
Stainless foil wrap is available in the following grades and thick-
nesses:
• For general-purpose, light-duty heat treating, use regular foil,
type 321 stainless, 0.002-in. (0.05-mm) thick for temperatures
up to 2,000° F (1,093° C).
• For heavy-duty, large, heavy, or sharp sections that may rip,
tear, or pierce the foil, use heavy-duty, type 321 stainless,
0.003-in. (0.08-mm) thick for temperatures up to 2,000° F
(1,093° C).
• For steels like high speed with high austenitizing tempera-
tures, use high-temperature foil, type 309 stainless, 0.002-in.
(0.05-mm) thick for temperatures up to 2,240° F (1,227° C).

GAS ATMOSPHERES
When tool steels are heated to temperatures above 1,000° F (538°
C), their surfaces become susceptible to chemical changes, depend-
ing on the furnace atmosphere. If the steel is subjected to an oxi-
dizing atmosphere (from oxygen in the air that leaks into the
furnace chamber), the surface carbon of the steel will combine
with the oxygen in the furnace, resulting in decarburization of
the steel. If the furnace atmosphere is reducing (that is, with a
high carbon potential and without oxygen), carbon will be infused
into the steel’s surface, resulting in carburization.
For most applications, the surface carbon content of the steel
being treated should not be altered from its basic carbon content;
that is, it should be neither carburized nor decarburized. There-
fore, tool steels should be heat treated in furnaces with atmo-
spheres that are not oxidizing or reducing, but neutral for the
grade of steel being treated.
The most common atmospheres for the neutral hardening of
tool steels include: inert gases, endothermic, exothermic, and
vacuum.

159

Ch08.pmd 159 10/29/02, 11:24 AM


Tool and Die Making Troubleshooter

Iner t Gases
The use of inert gases in the furnace atmosphere (like argon or
nitrogen) is the simplest way to avoid scaling and decarburiza-
tion. When an inert gas is pumped into a gas-tight furnace cham-
ber, it will not react with the hot steel and decarburization can be
avoided. Unfortunately, both of these gases are relatively expen-
sive and, for this reason, see limited use.

Endothermic Atmospheres
Endothermic atmospheres are formed when a mixture of air
and natural gas or propane is passed over a heated catalyst. The
heat and the catalyst make the gas and air react to form a mix-
ture of nitrogen (40%), carbon monoxide (20%), and hydrogen
(40%), plus some residual water vapor. After heating and mixing,
the gas mixture is cooled. The endothermic gas thus formed is
then pumped into the furnace for neutral hardening. The carbon
potential of the endothermic gas is a function of the dew point
(amount of moisture) in the gas atmosphere. The dew point must
be adjusted for each grade of tool steel being treated. Flow rates
of the gases into the furnace, as well as positive pressure in the
furnace, must be properly maintained. Done properly, carburiza-
tion or decarburization can be controlled.

Exothermic Atmospheres
Exothermic furnace atmospheres are the simplest and least
expensive. This atmosphere is produced through the partial com-
bustion of an air-gas mixture (frequently natural gas or propane
is used). This mixture is then burned in the presence of a catalyst.
Moisture formed during the combustion must be controlled as it
condenses in a cooling chamber. Exothermic atmosphere quality
is dependent upon using a very uniform composition of supply
gas prior to mixing with air and passing through the catalyst.

Cautions
Both endothermic and exothermic gases will burn when mixed
with air and, under certain conditions, can be explosive. There-

160

Ch08.pmd 160 10/29/02, 11:24 AM


Chapter 8: Heat Treatment

fore, when starting up a furnace, do not introduce the atmospheric


gases until the furnace temperature is up to or exceeds 1,400° F
(760° C) (Haga 1989a, b, 1974).

Vacuum
A vacuum atmosphere is the best way to protect steel surfaces
in heat treatment. Vacuum furnaces are electrically heated. Us-
ing a vacuum pump and a tightly sealed heat chamber, the steels
are heat treated in a near-vacuum. Thus, oxygen levels are re-
duced low enough so that they will not affect the steel when it is
heated to elevated temperatures, and carburization and/or decar-
burization cannot take place. A typical vacuum furnace is depicted
in Figure 8-7.

VACUUM HEAT TREATMENT


Protecting the tool surface is unquestionably one of the most
important steps in the heat-treating procedure. Hot steel will

Figure 8-7. Typical vacuum furnace (Kimble 1977).

161

Ch08.pmd 161 10/29/02, 11:24 AM


Tool and Die Making Troubleshooter

chemically react with whatever is in its atmosphere, whether it is


liquid or gaseous. Oxygen reacts with steel to form scale on the
surface and decarburization below the scale. Other gases react to
add carbon (carburization). Properly done, vacuum-furnace pro-
cessing is the best method for protecting tool-steel surfaces dur-
ing heat treating. The result is a bright-surfaced tool, free of
carburization or decarburization.
The word “vacuum” can be misleading. The “perfect” vacuum
has yet to be achieved, and it is impossible to completely empty
the atmosphere from a furnace chamber, even with the most so-
phisticated vacuum-pumping equipment. Therefore, the measure
of vacuum is really a measure of what is left in the chamber after
most of its original contents, usually air, is pumped out. Any quan-
tity of gas confined in a chamber exerts some pressure that is a
measure of the amount of gas remaining.
Units for measuring vacuum are very small: inches or microns
(µm or 0.001 mm) of mercury. A joint-research venture between
Bethlehem Steel Corp. and Lindberg Corp. concluded that vacuum
heat treatment for most tool steels can be conducted satisfacto-
rily in a range of 0.39–0.78 in. (100–200 µm) of mercury. However,
the furnace must be in good operating condition and the leak rate
must be very low.
From a safety standpoint, a vacuum of 0.20–0.39 in. (50–100
µm) of mercury or lower should be used if possible to ensure against
decarburization resulting from leakage or moisture of the inert
gas used for the quench. All work fixtures, baskets, etc., going
into the furnace must be scrupulously clean prior to treating
(Sullivan 1974).

Basic Principles
Although a vacuum furnace may seem a complex piece of pro-
cessing equipment, it is really a simple system. The operator must
follow the operational guidelines set by the equipment manufac-
turer and avoid furnace overloading.
Proper loading of a vacuum furnace is critical to its performance
and how the steel will respond. Parts must be spaced far enough
apart to allow for adequate gas circulation. Too dense a load can
cause the following problems:

162

Ch08.pmd 162 10/29/02, 11:24 AM


Chapter 8: Heat Treatment

• a long heat-up rate to the point where, depending on load


mass, the load does not achieve the proper hardening tem-
perature;
• a “slack quench” later when pressure quenching due to a
shielding effect, leading to incomplete metallurgical trans-
formation and soft working;
• a nonuniform gas-quench stage, leading to distortion and
possible shrinkage; and/or
• temperature irregularities that can cause nonuniform trans-
formation and irregular metallurgical results.
Placement of tooling in a vacuum-furnace work basket prior to
loading into the chamber is extremely critical, especially if the
cross sections of the parts vary. If parts must be mixed with dif-
ferent geometrical cross-sectional areas, try to keep the thinner
sections together, allowing clearance between parts so that gas
can circulate.

Quenching
Some vacuum furnaces employ built-in quenching chambers
kept under relative vacuum. Parts are immersed in the liquid (usu-
ally oil) after they have been properly austenitized. Other vacuum
furnaces employ a backfill quenching system that rapidly recircu-
lates an inert gas (usually nitrogen or argon) in the hardening
chamber after vacuum pumping has been stopped. The speed of
backfill quenching can be very rapid, approaching 1,000° F (538°
C) per minute. This rate is more than fast enough to thoroughly
quench air-hardening and many oil-hardening steels.

ANNEALING ALLOY AND TOOL STEELS


Annealing of steel reduces hardness, improves machinability,
facilitates cold working, and produces a desired microstructure.
Most alloy and tool steels are supplied in the annealed condition.
However, situations arise when the steels have to be annealed by
an end user, such as when material is purchased in the as-forged
or as-rolled condition, or prior to rehardening an already hard-
ened tool or one in preparation for weld repair.

163

Ch08.pmd 163 10/29/02, 11:24 AM


Tool and Die Making Troubleshooter

Slow heating to the proper annealing temperature for the grade,


adequate soaking at the designated high-temperature range, and,
above all, very slow, uniform, controlled cooling cycles are impor-
tant steps in a successful annealing operation. Table 8-4 presents
typical anneal cycles for various tooling steels.
Some annealing tips to remember:
• Anneal when an optimum, machinable structure before hard-
ening is needed.
• Anneal when steels must be rehardened, especially if they
are 45 HRC or higher.
• Anneal when distortion and size changes make it necessary
to rework tooling.
• Use a controlled-atmosphere, neutral salt, or vacuum furnace.
• Heat slowly and uniformly to the proper annealing tempera-
ture.
• Carefully monitor the furnace and part temperature.
• Slow cooling is a critical part of every annealing cycle.

NORMALIZING
Normalizing is a thermal treatment that heats steel to a suit-
able temperature, usually 100° F (56° C) above the steel’s trans-
formation range (that is, the upper critical temperature on
heating), followed by air cooling to ambient temperature. Low-,
medium- and high-carbon steels, some alloy steels, some stainless
steels, and some cast weldments are frequently normalized. Nor-
malizing for these steels is used as: (1) a hardening or strengthen-
ing heat treatment and (2) a preliminary conditioning procedure,
prior to annealing or hardening, to refine the structure of the
material being processed. However, normalizing is not applicable
to most tool steels, because they will harden when air cooled from
above their transformation range. Table 8-5 presents typical nor-
malizing temperatures for some popular carbon and alloy steels.
Years ago, normalizing was the principle hardening or strength-
ening method for large sections, such as large-diameter shafting
for heavy industry. This was because liquid quenching was con-
sidered too hazardous from a cracking standpoint. However, steel-
making advances—particularly vacuum degassing—made liquid

164

Ch08.pmd 164 10/29/02, 11:24 AM


Ch08.pmd

Table 8-4. Typical anneal cycles for various tooling steels


AISI Anneal, Soak Cycle, Slow
Slow--cool PProcess,
rocess, Brinell
Grade °F (°C) min/in. (min/mm) °F/hr (°C/hr) Hardness
W1, W2 1,400 (760) 30 (1.18) 50–975 (10–524) 200
O1 1,450 (788) Soak thoroughly 25–900 (–4 to 482) 221
O6 1,450 (788) Soak thoroughly 20–900 (–7 to 482) 220
165

S5 1,450 (788) 60 (2.36) 25–900 (4–482) 230


S7 1,550 (843) 90 (3.54) 25–900 (4–482) 230
A6 1,400 (760) 60 (2.36) 25–900 (4–482) 230
A2 1,650 (899) 120 (4.72) 40–900 (4–482) 235
165

A-HT 1,650 (899) 120 (4.72) 40–900 (4–482) 235


D2 1,650 (899) 90 (3.54) 20–900 (–7 to 482) 220
H13 1,600 (871) 60 (2.36) 25–900 (–4 to 482) 220
M2 1,600 (871) Soak thoroughly 25–900 (–4 to 482) 241
M4 1,600 (871) Soak thoroughly 25–900 (–4 to 482) 241

Chapter 8: Heat Treatment


10/29/02, 11:24 AM

420 SS 1,650 (899) Soak thoroughly Furnace cool to 600 (316) 230
4140 1,500 (816) Soak thoroughly 20–1,230 (–7 to 666) 215
(Bethlehem Steel Corp. 1976, 1977, 1978a–g, 1980, 1981a, Undated a, b; Teledyne Vasco 1968)
Tool and Die Making Troubleshooter

Table 8-5. Normalizing temperatures for carbon and alloy steels


Grade °F (°C)
1020 1,700 (927)
1060 1,650 (899)
4140 1,600 (871)
4340 1,600 (871)
6150 1,600 (871)
8620 1,700 (927)
(Bethlehem Steel Corp. 1980)

quenching of large sections less hazardous. As a result, normaliz-


ing is no longer as popular for hardening or strengthening large
steel sections. Normalizing for strength was commonly followed
by tempering, both to minimize residual stress and control me-
chanical properties to within desired limits.
The other use for normalizing is as a homogenizing treatment
(prior to annealing or hardening by quenching) for carbon and
alloy steels that are in the as-forged or as-rolled condition. Many
steels in either of these conditions tend to be relatively nonhomo-
geneous with respect to microstructure, and respond poorly to
annealing and hardening operations. The high normalizing tem-
perature erases this undesired microstructure, and the following
air-cool helps to develop a relatively fine, homogeneous micro-
structure that responds well to further thermal treatment.
Normalizing for those steels previously noted also improves
machinability, grain structure, grain refinement, and structural
homogeneity. Normalizing also is used to relieve steels and
weldments of residual stresses.

STRESS RELIEVING
Stress relieving steels prior to hardening is frequently a good
practice. If done properly, a large percentage of the mechanical
stresses imparted to the steel from cold working, saw cutting,
machining, etc., can be removed. Stress relieving after heavy ma-
chining helps to minimize problems related to the size change
and distortion of machine parts, weldments, and tools that often

166

Ch08.pmd 166 10/29/02, 11:24 AM


Chapter 8: Heat Treatment

develop upon hardening. The long, thin die components in Figure


8-8 had to be scrapped due to excessive distortion after being hard-
ened. Replacement parts that were stress relieved after being
machined and before hardening were finished to dimension with-
out difficulty.

Figure 8-8. D2 die sections scrapped due to excessive distortion after being
hardened.

Here is a basic stress-relief procedure for annealed alloy, tool,


and high-speed steels:
1. Heat slowly and uniformly to 1,200–1,250° F (649–677° C).
2. Soak at heat for 1–2 hours per 1 in. (25.4 mm) of thickness.
3. Slow cool in the furnace to room temperature.
Be careful to heat and cool uniformly, especially when a part
being stress relieved has a complicated design. If the rate of cool-
ing is not uniform, new stresses can develop, which could be equal
to or even greater than those that the stress-relief cycle was in-
tended to remove.
Stress relieving is becoming more important with the trend to-
ward taking faster and heavier machine cuts, employing compli-
cated design configurations, and machining closer to finished
dimensions before hardening.
Stresses imparted to tooling during machining and heat treat-
ment may combine to cause severe distortion. This dangerous
combination of stresses frequently causes bending, bowing, twist-
ing, and/or cracking of the steel. The procedure for stress reliev-
ing tool steels is simple. Problems related to stress relief are most
often due to shortcuts taken to reduce the overall time needed to
heat and cool the steel uniformly and slowly.

167

Ch08.pmd 167 10/29/02, 11:24 AM


Tool and Die Making Troubleshooter

After stress relieving, it is important that the tools be remea-


sured. Any dimensions that have changed, as a result of the stress
relief operation, can be corrected with additional light machin-
ing. Plan ahead in the rough machining operation and leave suffi-
cient stock for correction to be made after stress relieving.

PREHEAT BEFORE HARDENING


Almost every publication on tool steels recommends that tools
be carefully preheated immediately before being heated to the
quenching temperature. If no appreciable amount of cold work
has been done, preheating is not actually necessary if the quench-
ing temperature is below 1,600° F (871° C). However, when the
quenching temperature is 1,600° F (871° C) or higher, it is advan-
tageous to preheat tools to prevent thermal shock, provide uni-
formity of heating to the hardening temperature, minimize scaling
and decarburization, and promote dimensional stability. Preheat-
ing also should be employed if a tool has been appreciably cold
worked, if it is not in the fully annealed condition, or if it has a
complicated design with numerous stress raisers. Table 8-6 pre-
sents typical preheating cycles for selected tool and high-speed
steels.

Table 8-6. Typical preheating cycles for tool and high-speed steels
Grade Temperature °F (°C) Soak Time
W1, O1, O6 1,200–1,250 (649–677) Thorough at heat
A2, A6, A-HT 1,200–1,250 (649–677) Thorough at heat
D2, D3 1,200–1,250 (649–677) Thorough at heat
S1, S2, S5, S7 1,200–1,250 (649–677) Thorough at heat
H13 1,300–1,400 (704–760) Thorough at heat
H21 1,300–1,500 (704–816) Thorough at heat
M2, M4 High-speed Steels: Two-stage Preheat Cycle
1st stage 1,550 (843) Thorough at heat
2nd stage 1,850 (1,010) Thorough at heat
(Bethlehem Steel Corp. 1954, 1976, 1977, 1978a–f, 1981a, Undated a, b, c;
Teledyne Vasco 1968)

168

Ch08.pmd 168 10/29/02, 11:24 AM


Chapter 8: Heat Treatment

Preheating prior to hardening will help to:


• reduce mechanical stresses (from cold working) and thermal
stresses (from heating and cooling) that can cause warpage
and/or cracking during hardening;
• minimize scaling and decarburization of the workpiece;
• promote temperature uniformity during transformation on
subsequent heating and quenching;
• reduce the time the workpiece is exposed to high austenitizing
temperatures; and
• improve dimensional stability (resistance to bending, bow-
ing, and/or twisting) of the workpiece during the hardening
operation.
Note that the benefits of preheating become more important as
the design of the workpiece becomes more complicated with stress
raisers—sharp corners, dramatic changes in cross section, thick
sections adjacent to thin sections, numerous holes, thin wall sec-
tions, large and/or long thin sections, etc.

HARDNESS-CAUSED CRACKING
Heat treating tooling components to maximum hardness levels
is a frequent cause of premature tool-and-die failure. The belief
that the higher the hardness, the longer and better a part will
perform in service, is an oversimplification. For most tooling ap-
plications, wear resistance and toughness requirements must be
considered and a hardness range selected to produce these prop-
erties.
Figure 8-9 shows an S7 tool-steel punch that cracked after very
light service. This part was hardened and tempered to 60–61 HRC.
Hardening this punch to 56–58 HRC would have dramatically in-
creased its service life as the toughness of S7 drops significantly
with hardness higher than 58 HRC.
Keep in mind that as the hardness of a steel increases, its over-
all toughness and ductility will decrease. To develop maximum
hardness levels, heat treaters are often required to increase hard-
ening temperatures, lengthen soak times at the hardening tem-
perature, and increase the speed of the quench. After hardening,

169

Ch08.pmd 169 10/29/02, 11:24 AM


Tool and Die Making Troubleshooter

Figure 8-9. S7 tool-steel punch that cracked after light service.

they may be required to short cycle tempering to avoid temper


softening. Any one of these adjustments to secure higher hard-
ness may result in brittle parts. Alone or in combination, they
almost always result in premature service failures from chipping
and cracking.
Whenever heat-treat specifications call for hardness levels that
exceed normal working hardness ranges, a flag of caution should
be raised. The entire hardening process should be reviewed. Care-
ful processing should be followed to ensure an optimum balance
of hardness, toughness, and minimal residual stresses. Table 8-7
lists the maximum and normal working hardness for many of the
popular tool steels used today.
The M2 high-speed-steel punch in Figure 8-10 cracked after
very light service. This part was hardened and tempered to 63–65
HRC for maximum wear resistance. A lower hardness in the range
of 60–62 HRC would have dramatically increased its ductility,
toughness, and production performance.

170

Ch08.pmd 170 10/29/02, 11:24 AM


Chapter 8: Heat Treatment

Table 8-7. Working hardness for some popular tool steels


Maximum Normal Recommended
AISI Obtainable Working Tempering
Tool Steel HRC* HRC Temperature, °F (°C)
A2 64 58–60 400–450 (204–232)
D2 64 58–60 900–960 (482–516)
H13 54 40–48 1,050–1,150 (566–621)
M2 65 62–64 1,000–1,050 (538–566)
O1 64 58–60 400–450 (204–232)
S7 60 56–58 400–450 (204–232)
* as-quenched hardness
(Bethlehem Steel Corp. 1976, 1977, 1978d-f, 1981a)

Figure 8-10. M2 high-speed punch that cracked after light service.

171

Ch08.pmd 171 10/29/02, 11:24 AM


Tool and Die Making Troubleshooter

AUSTENITIZING SOAKING TIMES


Tool-steel manufacturers have established the austenitizing
(hardening) temperature that will attain maximum hardness and
grain refinement for each grade of tool steel and the holding times
required for austenitization and alloy/carbide solution. Table 8-8
shows the recommended austenitizing cycles for some selected steels.

Table 8-8. Recommended austenitizing soak cycles


Minimum Soak Times,
Grade Temperature, °F (°C) min/in. (min/mm)
W1, W2 1,400–1,450 (760–788) 30 (1.18)
O1 1,475–1,500 (802–816) 30 (1.18)
O6 1,450–1,500 (788–816) 45–60 (1.77–2.36)
A2 1,750–1,800 (954–982) 45–60 (1.77–2.36)
A6 1,550 (843) 45–60 (1.77–2.36)
S5 1,600 (871) 30 (1.18)
S7 1,725 (941) 45–60 (1.77–2.36)
H13, D2 1,850 (1,010) 45–60 (1.77–2.36)
420 SS 1,750–1,850 (954–1,010) 45 (1.77)
4140 1,550 (843) 30 (1.18)
M2, M4 2,150–2,250 (1,177–1,232) 5–15 (0.19–0.59)
(Bethlehem Steel Corp. 1976, 1977, 1978a–g, 1980, 1981a, Undated b; Teledyne
Vasco 1968)

The manner in which the recommended soaking cycle is attained


is often a matter of experience. Variations in heat-treating tech-
nique, equipment, as well as the size, shape, and mass of the tool-
ing make it difficult for the heat treater to follow a fixed set of
thermal-cycling instructions. Heat treaters must frequently bal-
ance, modify, and blend the instructions and procedures recom-
mended by the tool-steel producer with the capabilities and/or
limitations of the processing equipment. The heat treater also must
be able to define a proper soak cycle for every tooling situation
that develops. The following tips should be carefully reviewed:

172

Ch08.pmd 172 10/29/02, 11:24 AM


Chapter 8: Heat Treatment

• Soak cycles must be long enough to ensure temperature uni-


formity of the steel at hardening temperature and allow for
proper alloy and carbide solution.
• Excessive grain growth and austenite retention will result
from oversoaking or a hardening temperature that is too high.
• If the hardening temperature is considerably below that rec-
ommended, hardening will not occur. A somewhat higher tem-
perature (yet, still below the recommended) may result in
nonuniform hardening, with the surface capable of attaining
hardness, but the interior ineffectively hardened. This can
produce quench cracking.
• Heating at temperatures exceeding the normal hardening
temperature results in grain coarsening and alteration of the
microstructure with inherent brittleness. Quenching from
these temperatures is more likely to produce cracking.
• Lack of sufficient soak time during austenitizing can result
in insufficient alloy and carbide solution in the higher-alloy
tool steels, and nonuniform hardening. For parts under 1 in.
(25.4 mm) in thickness, a minimum soak of 45–60 minutes
at heat is normally required.
• Excessive holding time can promote grain coarsening and lead
to early tool failure.

QUENCHING
Quenching is rapid cooling from the austenitizing (hardening)
temperature. It may be accomplished in many ways, for example,
via oil, water, brine, still air, or salt baths, depending on the mate-
rial being heat treated.
The quenching of tool-and-die steels is the most critical step in
the heat-treatment process. It is during the quenching cycle that
hardness and accompanying physical properties are developed. In
general, tool steels are quenched at different rates depending on
their alloy content, that is, in water, oil, or air. In comparison,
water-hardening steels have little or no alloy content, oil-harden-
ing steels have a higher alloy content, and air-hardening steels
are considered to be relatively high in alloy content (Bethlehem
Steel Corp. 1981b).

173

Ch08.pmd 173 10/29/02, 11:24 AM


Tool and Die Making Troubleshooter

Rates and Stresses


Water quenching is the fastest of the quench cycles, producing
the greatest amount of stresses. Oil quenching is approximately
one-third slower than water quenching, and air quenching is very
slow by comparison. To ensure the development of hardness, a
carbon tool steel should be quenched below approximately 1,000°
F (538° C) in less than 1 second. Oil-hardening tool steel needs to
be quenched below about the same temperature in 8–10 seconds
maximum. D2, an air-hardening steel, can be cooled to below a
temperature of about 1,300° F (704° C) for as long as 4 minutes
and still achieve full hardness.
When austenite is transformed to martensite during the quench-
ing operation, the transformation is accompanied by a physical
change in the size of the material being heat treated. Martensite
occupies more volume than austenite. This increase in volume
results in a significant increase in the residual stress level of the
part being heat treated. In addition, cooling of the material from
the high hardening temperature to the low temperature at the
end of the quenching operation results in the formation of stresses
that increase with the rate of cooling.
The net increase in the stress level—which results from a com-
bination of the cooling and transformation stresses—may well
approach the ultimate strength level of the steel. If a stress raiser—
such as a sharp corner or a thin section adjacent to a heavy sec-
tion—is present, the stress level may locally exceed the strength
of the material and cracking will result. Therefore, it is important
to pay attention to tool design and eliminate obvious stress rais-
ers. If this is not possible, then an air-hardening steel must be
selected for the part.
Air-hardening tools should be quenched to 150° F (66° C). Oil-
hardening and water-hardening tools should be quenched to about
150–200° F (66–93° C). The recommended temperatures for
quenching baths are as follows: water, maintain at 70–90° F (21–
32° C); and oil, 90–130° F (32–54° C). Bath agitation in a liquid
quench is required to allow proper heat transfer from the parts
being quenched. Air hardening can usually be performed in still
air, depending on the section size.
Here are some common concerns about quenching of tool-and-
die steels during the hardening operation:

174

Ch08.pmd 174 10/29/02, 11:24 AM


Chapter 8: Heat Treatment

• Do not quench the steel being hardened until it has been ad-
equately soaked at the hardening (austenitizing) temperature.
• Temper immediately after quenching.
• Do not allow as-quenched steels to sit at ambient tempera-
tures for long periods of time before tempering.
• Strive to quench the workpiece or batch as uniformly as pos-
sible.
• Quench parts (especially long, thin sections) vertically to mini-
mize concerns relating to bending, bowing, and twisting.

Water Quenching
For water quenching, brine is preferred over fresh water. This
is because the steam formed when the quench water contacts the
superheated part forms a vapor barrier that insulates the parts
being quenched from proper heat transfer, particularly at sharp
internal corners, machined threads, blind holes, rough machine
marks, and other recesses. The result is the development of soft
spots on the parts being quenched and/or the development of dif-
ferential quenching and transformation stresses that may result
in distortion and/or cracking.
The addition of brine (up to 10% maximum by volume) facili-
tates the quenching process because salt crystals that precipitate
on the surface of the part being quenched violently explode away.
This explosive action causes a severe agitation that prevents the
formation of vapor barriers on the parts being quenched. The agi-
tation also throws off much of the heat-treat scale and makes the
quench action more uniform. Agitating the quenchant or the work-
piece prevents the formation of vapor barriers and facilitates more
uniform cooling.

Oil Quenching
Here are some tips on oil quenching:
• Because quenching in oil can be a fire hazard, the use of
quench oils with high flash points is recommended.
• Oil quenching is slower that water or brine quenching,
thereby causing lower residual quenching stresses than ei-
ther of the latter.

175

Ch08.pmd 175 10/29/02, 11:24 AM


Tool and Die Making Troubleshooter

• It takes approximately 1 gal of oil for each 1 lb (1.7 L per kg)


of steel quenched per hour. If 100 lb (45 kg) of steel is hard-
ened every hour, for example, a tank of 100-gal (378-L) ca-
pacity is required.
• Oil quenchants should be kept at a temperature of 90–130° F
(32–54° C) to provide proper quenching action.
• Agitate the quenchant or the workpiece to facilitate uniform
cooling.

Air Quenching
For quenching in air, add these considerations:
• Cool the workpiece as uniformly as possible during a still-air
quench.
• Fan-air blasting may be used to promote uniformity or accel-
erate cooling.

Integral Vacuum-furnace Quenching


Vacuum heat treaters should keep in mind that the speed of the
internal quench in the vacuum furnace must be carefully moni-
tored and adjusted to suit the steel being processed and the load
conditions.
Quench rates that are too slow and/or nonuniform may pro-
duce lower than required hardness levels. Quenching too fast
can cause cracking of parts with thick or massive cross sections,
thin and thick sections, and other types of stress raisers.

Quench Cracking
Quench cracking is the most common cause of failure in heat
treatment. Quench cracks develop when stresses introduced into
the steel combine to equal or exceed its ultimate strength. From a
stress-development standpoint, water quenching is the fastest of
the quench cycles, but induces the highest amount of stress. Oil
quenching is approximately one-third slower than water quench-
ing and correspondingly lower in terms of stress generation. Air
quenching is comparatively slower and less stressful than oil

176

Ch08.pmd 176 10/29/02, 11:24 AM


Chapter 8: Heat Treatment

quenching. The H13 tool-steel hot extrusion die shown in Figure


8-11 was one of six that cracked as a result of a poor design for a
liquid-quenching steel. Failure to temper immediately after
completion of the quench cycle was also a contributing factor.
Two important precautions must be observed if quench crack-
ing is to be minimized or eliminated:
• Remove tools and dies from the quench cycle while they are
still warm. Water- and oil-hardened tools should be quenched
to approximately 150–200° F (66–93° C) and tempered im-
mediately. Air-hardening tools and dies should be quenched
no colder than 150° F (66° C) before being immediately tem-
pered. It is also important that the baths for oil and water
quenching be maintained as follows: water at 70–90° F (21–
32° C), and oil at 90–130° F (32–54° C).
• Transfer the as-quenched tools immediately to a preheated
tempering furnace. Placing tools in a preheated tempering
furnace immediately after reaching the proper quenching

Figure 8-11. Cracked H13 tool-steel extrusion die.

177

Ch08.pmd 177 10/29/02, 11:25 AM


Tool and Die Making Troubleshooter

temperature for the particular grade of steel is important to


relieve stresses that develop during quenching. These stresses
may be close to the ultimate strength of the steel. If they
develop to the point where they equal or exceed the ultimate
strength of the steel—or if stress raisers cause concentra-
tion at a particular location to the point where the stresses
equal or exceed the ultimate strength of the steel—cracking
will occur.
Note: Careful handling in the quench and immediate temper-
ing processes may not resolve all quench cracking problems. Stress
concentrations and conditions that cause formation of the differ-
ential transformation stresses may cause cracking to occur no
matter how carefully the quenching operation is conducted. Some
of these are:
• There is failure to maintain surface chemistry during hard-
ening (particularly the carbon content). The development of
carburization or decarburization during austenitizing often
results in cracking during the quench because of the differ-
ential transformation stresses that develop between the sur-
face metal (either carburized or decarburized) and the parent
metal.
• Design faults—such as sharp corners, improperly spaced
holes, thin sections adjacent to heavy sections, etc.—all act
as stress raisers. If these stress raisers cannot be avoided in
a design, it is suggested that air-hardening steels be used.
• Quenching a tool from a much higher austenitizing tempera-
ture than recommended is dangerous. This practice results
in grain coarsening of the steel, which makes it brittle, and
the development of higher quenching stress than would de-
velop with the lower, correct austenitizing temperature.
• Quenching a tool from a temperature below that recom-
mended for the grade and/or insufficient soaking at the
hardening temperature is another dangerous practice. Non-
uniform hardening—with the surface capable of attaining
hardness, but the interior being only partially hardened—
can result in the development of differential transformation
stresses, which contribute to quench cracking.

178

Ch08.pmd 178 10/29/02, 11:25 AM


Chapter 8: Heat Treatment

TEMPERING
During the hardening process, tooling must be tempered im-
mediately following quenching to relieve residual quenching
stresses that can be the cause of cracking. Tempering also is per-
formed to achieve specific hardness levels, and especially in the
case of air-hardening steels, to promote transformation of any
austenite retained after quenching to martensite. To get the maxi-
mum benefit of tempering, the operation should never be short
cycled. Figure 8-12 presents the tempering versus hardness curves
for American Iron and Steel Institute (AISI) D2 and S7 tool steels.
Good tempering practice relieves stresses that develop during
quenching, aids the transformation of retained austenite (espe-
cially in air-hardening steels upon cooling from the tempering
process), and improves the dimensional stability of tooling. Here
are some tips for proper tempering:

Figure 8-12. Tempering curve versus hardness for AISI D2 and S7 tool
steels (Bethlehem Steel Corp. 1976, 1978d).

179

Ch08.pmd 179 10/29/02, 11:25 AM


Tool and Die Making Troubleshooter

• Never allow as-quenched tooling to be cooled to room tem-


perature.
• Air-hardening tool steels should be cooled to a minimum of
150° F (66° C). Liquid quenched tools should be cooled to
150–200° F (66–93° C). Then both should be tempered imme-
diately.
• The tempering furnace should be preheated.
• Soaking times for the tempering should not be less than 120
min/in. (4.7 min/mm) of thickness at heat for temperatures
up to 1,000° F (538° C), and 60 min/in. (2.4 min/mm) of thick-
ness at heat for steels tempered at 1,000° F (538° C) or higher.
• High-carbon, high-chromium, air-hardening, and high-speed
tool steels should always be double tempered, and sometimes
triple tempered. Whenever possible, temper these steels in the
range of secondary hardness (note the curve for D2 in Figure
8-12), as these temperatures are best for promoting trans-
formation of retained austenite to martensite upon cooling
from the temper.
The true benefits of proper tempering after hardening should
never be underestimated. There is no substitute for proper plan-
ning, especially to eliminate the time restraints and scheduling
pressures that frequently result in short cycling of the tempering
operation. Remember that although some of the benefits of tem-
pering, such as stress relief and thorough transformation, are in-
visible to the naked eye, they are just as important as resulting
hardness ranges in terms of the service performance of tools and
dies.
The D2 tool-steel die in Figure 8-13 cracked after very little
service. Review of the heat-treating procedure revealed that it
was tempered well below the range of secondary hardness (925–
950° F [496–510° C]) and for too short a time.

SUBZERO TREATMENTS
Subzero treatments have been used to improve the quality of
tooling manufactured from steels that have a tendency to retain
austenite. The addition of deep cryogenic tempering (approxi-
mately –300° F [–184° C]) to subzero treating operations can sig-

180

Ch08.pmd 180 10/29/02, 11:25 AM


Chapter 8: Heat Treatment

Figure 8-13. AISI D2 tool-steel die, 7.50 in. (190.5 mm) in diameter with
crack radiating across the face and down the outside diameter.

nificantly improve tool performance on virtually all grades of tool-


ing steels. Deep cryogenic tempering reportedly does the follow-
ing: (1) creates a denser molecular structure, resulting in a larger
contact area that reduces friction, heat, and wear; (2) decreases
residual stress; and (3) increases tensile strength, toughness, and
dimensional stability.

181

Ch08.pmd 181 10/29/02, 11:25 AM


Tool and Die Making Troubleshooter

The proper time for subzero treatment is after an initial tem-


per. Never subzero treat immediately after the quench. Figure 8-
14 shows the fundamental heat-treating steps required for tool
steels. Note the deep freeze and cryogenic options after the first
temper.

Figure 8-14. Fundamental heat-treating steps for tool steels.

Benefits and Procedures


The benefits to be gained from subzero treating can be summa-
rized as:
• increased wear resistance because of maximum transforma-
tion of austenite;
• better mechanical properties, hardness, and toughness;
• increased dimensional stability;
• salvaging of tooling made soft by austenite retention; and
• reduced cracking hazards during subsequent EDM, grind-
ing, and welding.
The step-by-step procedure is as follows:
1. Use normal hardening procedures.
2. Temper before subzero treatments to stress relieve.
3. For a –120° F (–84° C) deep freeze, soak 120 min/in. (4.72
min/mm) of greatest thickness.
4. For cryogenics, do not plunge hardened tooling directly into
liquid nitrogen. Controlled slow cooling to –320° F (–196° C)
is safest.

182

Ch08.pmd 182 10/29/02, 11:25 AM


Chapter 8: Heat Treatment

5. A minimum 24-hour cycle for cryogenic tempering is recom-


mended. Consult a heat-treating specialist for specific proce-
dures.
6. Slowly return parts to room temperature.
7. Retemper immediately after all forms of deep freezing and
cryogenic tempering (Sweeney 1986; 300° F Below, Inc. Un-
dated).

DISTORTION AND SIZE CHANGE


Is it worth the gamble to machine tools and dies to close finish
dimensions before hardening? This has been an increasing trend
to minimize the need for finish machining (such as grinding and
EDM) hardened steels and lower machining costs. Unfortunately,
this is a dangerous practice because dimensional changes can oc-
cur during hardening that frequently do not work out as planned.
If sufficient stock is not left on a tool or die prior to heat treat-
ment, the resultant combination of size change and/or distortion
may make it impossible to finish-machine the part to required
dimensions.
It is important to understand the sources of stresses that can
result in size change and distortion of tools and dies in heat treat-
ment. First, note that nondeforming or distortion-free tool steels
do not exist. All tool steels are subject to some warpage (bending,
bowing, twisting) in combination with growth or shrinkage when
hardened. The degree of distortion depends on the amount of re-
sidual stresses that develop in the steel. Volume changes not only
result in dimensional changes (growth or shrinkage) of the tool or
die, but also increase the residual stress level.
The basic causes of size change and distortion are: mechani-
cally induced stress, thermal stress, transformation stress, and
combination stresses.

Mechanically Induced Stress


Mechanically induced stresses result from cold-working opera-
tions (machining, saw cutting, etc.). The magnitude of these
stresses is in direct proportion to depth of cut, cutting speed, etc.
If these stresses are not relieved before a tool or die is heat treated,

183

Ch08.pmd 183 10/29/02, 11:25 AM


Tool and Die Making Troubleshooter

they will be released during the heating for hardening and may
result in distortion (bending, bowing, and/or twisting).

Thermal Stress
Thermally induced stresses during heating and cooling also in-
crease the residual stress level of a part. When steel is heated to
the preheat and hardening temperature, it expands. The mass
and design configuration of the tool or die will determine how
uniformly the part will expand during heating. Light cross sec-
tions expand more rapidly and will reach temperature before heavy
sections. The surfaces of heavy cross sections will reach tempera-
ture before interior sections. The noted temperature differences
result in differential rates of expansion between light and heavy
sections, and the surface and interior sections. This differential
expansion will result in an increase of the residual stresses until
such time as the part is at uniform temperature throughout.
When the same part is cooled during the quenching cycle, the
reverse happens. Light cross sections and the surfaces of the
heavier sections of the tool or die cool faster than either the heavy
cross sections or the interior. These temperature differences also
result in differential contraction, which in turn results in addi-
tional residual stress development.

Transformation Stress
Transformation stresses, which develop on heating and cool-
ing, are also responsible for inducing residual stresses. When the
steel being heated for hardening reaches its critical temperature
on heating, the existing annealed microstructure transforms to
austenite. This transformation results in shrinkage of the part be-
cause austenite is smaller in volume than the annealed structure
from which it develops. After all of the annealed microstructure is
transformed to austenite on further heating to the hardening tem-
perature, the part begins to expand again.
All of the expansion and contraction discussed thus far will occur
at different rates in light cross sections as opposed to heavy cross
sections, or in surface locations as opposed to interior locations.
In this way, stress levels are compounded as part geometry ex-
pands and contracts at different times during heating and cooling.

184

Ch08.pmd 184 10/29/02, 11:25 AM


Chapter 8: Heat Treatment

Transformation also occurs on cooling during the quench. The


austenite formed on heating to the hardening temperature trans-
forms to martensite during the quench. Martensite is larger in
volume per unit of mass than either the austenite from which it
transformed or the annealed structure that existed before heat
treatment. Again, because the transformation takes place faster
in lighter cross sections and at surface locations as opposed to
interior locations, differential stresses will occur in the part.

Stress Combinations
The combination of residual stresses from machining and the
thermal and transformation stresses that occur during harden-
ing may result in the distortion of a tool or die. Such distortion
can be minimized by relieving machining stresses prior to hard-
ening and heating to the hardening temperature at a rate that
will result in minimal temperature differential in the part. While
stresses that occur on heating will be relieved when soaking at
the hardening temperature is accomplished, the distortion that
occurs may not be reversed. Likewise, the combination of ther-
mal and transformation stresses upon cooling (quenching) may
also result in distortion, which will not be removed by subsequent
tempering.

Volume Changes from Transformation


On quenching, the overall dimensions of the part will change
due to the growth during heat treatment. The amount of growth
has been found to be partly dependent on the steel type. Table 8-9
shows the expected growth of various types of tool steels in heat
treatment. Note that shrinkage may develop on a given dimension
if full transformation of austenite to martensite is not realized.
A word of caution: The values in Table 8-9 are not absolute.
Experience indicates that part size, mass, geometry, and steel type
play roles in dictating which dimensions of the part will grow in
heat treatment and which will shrink. For example, a long, thin
rectangular shape that hardens through completely in heat treat-
ment will grow in all dimensions (that is, width, thickness, and
length). However, a long, thick rectangular shape that does not

185

Ch08.pmd 185 10/29/02, 11:25 AM


Tool and Die Making Troubleshooter

Table 8-9. Growth of various types of tool steels in heat treatment


Expected Growth, in. per in.
Steel Type (mm per mm) of TTool
ool Dimensions
Water hardening (W1, W2, etc.) + 0.002–0.004
Oil hardening (O1, O6, etc.) + 0.0015
Air hardening (A2, A6, S7, H13, etc.) + 0.001
Air hardening (D2) ± 0.0005
(Bethlehem Steel Corp. 1965)

harden through will grow in width and thickness, but will shrink
in length.
Predicting just how much a steel tool or die will shrink, grow,
and/or distort is very difficult. Part design, mass, size, grade of
steel, and level of hardness all have an effect on stress levels and
the expected growth of a tool or die in heat treatment. All of these
factors complicate size predictability in heat treatment, making
it necessary to include adequate finish-machining allowance on a
part for removal after heat treatment. Insufficient or a zero fin-
ish-machining allowance will place a part in jeopardy, possibly
resulting in it being scrapped and having to be remade.
In addition to adding sufficient allowance for finish machining
after heat treatment, the use of air-hardening steels is recom-
mended, as these will result in minimal size change in heat treat-
ment. Stress relieving of tools before heat treatment and heating
at rates that minimize extreme temperature differentials between
thin and thick sections (as well as between the surface and inte-
rior of heavy sections) will help minimize distortion levels. Proper
support, especially during hardening and quenching, is also im-
portant. Tooling that can be suspended vertically, for example, is
much less likely to bend and bow during heat-treat processing
(Bethlehem Steel Corp. 1965, 1981a, Undated d).

TROUBLESHOOTING
Failure to achieve the required hardness, nonuniform surface
hardness, decarburization, scaling, distortion, and cracking during

186

Ch08.pmd 186 10/29/02, 11:25 AM


Chapter 8: Heat Treatment

heat treat are concerns that everyone involved in the construction,


application, and maintenance of tools, dies, molds, fixtures, and
machine parts has to address. This chapter has reviewed these
concerns as they relate to improper heat-treat practices and the
use of less-than-optimum equipment (assuming that the steel be-
ing processed is of good quality and within specification).
Most concerns relating to heat treatment result from a combi-
nation of poor planning and short cycling to save time. A proac-
tive approach that plans for and allows sufficient time to carry
out the necessary heat-treat cycles for preheating, soaking at the
hardening temperatures, quenching, and tempering, etc., will
eliminate a high percentage of these concerns.

Problem: High Surface Hardness


Possible causes and remedies:
• Poor furnace temperature control—check and calibrate fur-
nace controls regularly.
• Surface carburization—anneal and/or harden in furnaces
with controlled neutral atmospheres.
• Using hardening temperatures that are too high—do not ex-
ceed the manufacturer’s recommended hardening tempera-
ture range.
• Overheating and/or oversoaking at the hardening tempera-
ture—do not exceed the manufacturer’s recommended hard-
ening soak times, which vary with different furnace types.
• Insufficient and/or short cycling of the tempering operation—
select the highest tempering temperature consistent with the
required hardness. Soak at the tempering temperature for
120 min/in. (4.72 min/mm) of thickness at heat. Double and
triple temper air-hardening tool steels.

Problem: Low Surface Hardness


Possible causes and remedies:
• Poor furnace temperature control—check and calibrate fur-
nace controls regularly.
• Surface decarburization—anneal and/or harden in furnaces
with controlled neutral atmospheres.

187

Ch08.pmd 187 10/29/02, 11:25 AM


Tool and Die Making Troubleshooter

• Using a hardening temperature that is too low—stay within


the manufacturer’s recommended hardening temperature
range.
• Insufficient soaking at the hardening temperature—preheat
thoroughly before hardening. Follow recommended austen-
itizing (hardening) soak times.
• Quenching too slowly and/or quench interruptions—quench
uniformly, without interruption. Employ a uniform fan blast
for air-hardening steels when section size requires more than
a still-air quench for proper cooling.
• Tempering at too high a temperature range for the desired
hardness level—follow recommended tempering guidelines.

Problem: Nonuniform Surface Hardness


Possible causes and remedies:
• Surface decarburization and/or surface scale—be sure all
decarburization is removed prior to hardening. Anneal and/
or harden in controlled neutral atmospheres, or when these
are not available, use stainless foil wrap. Seal all furnace open-
ings to keep out air. Keep furnaces in good repair so they can
be properly sealed.
• Nonuniform quenching—when air quenching large sections,
use a uniform fan blast to accelerate cooling. Small section
sizes may be cooled in still air in a place where air circulation
is not impeded. Liquid quenchants should be agitated and
kept cool possibly with a chiller. Keep liquid-quench baths
clean. Use brine solutions instead of water quenchants.

Problem: Surface Decarburization and Scale


Possible causes and remedies:
• Failure to remove surface decarburization from barstock—
use decarburization-free steels whenever possible. Check and
remove surface scale and decarburization as recommended
from as-rolled stock.
• Failure to heat and soak in a neutral furnace atmosphere
during hardening—anneal and/or harden in a controlled neu-
tral atmosphere, vacuum, or neutral salt furnace.

188

Ch08.pmd 188 10/29/02, 11:25 AM


Chapter 8: Heat Treatment

Problem: Distor tion (Bending, Bowing, or Twisting)


Possible causes and remedies:
• Complicated design configurations—design part with mini-
mal stress raisers, avoiding sharp internal corner sections,
thick sections adjacent to thin sections, blind holes, and holes
with thin wall sections.
• Mechanical stresses from cold working (machining) were re-
leased—stress relieve prior to hardening to eliminate me-
chanical stresses.
• Insufficient part fixturing and support in the furnace—do
not overload furnaces. Support the workpiece properly in the
furnace, suspending it vertically whenever possible. Clamp
long, thin workpieces to support plates whenever possible.
• Thermal shock from heating to the hardening temperature
too rapidly—preheat thoroughly before hardening. Heat to
and soak uniformly at the hardening temperature.
• Failure to heat and soak uniformly at the hardening tem-
perature—preheat thoroughly. Do not overload furnaces. Soak
thoroughly at the hardening temperature.
• Nonuniform quenching—use air-hardening steels whenever
possible. Quench uniformly.

Problem: Dimensional Changes (Shrinkage or Growth)


Possible cause and remedy:
• Incomplete and/or nonuniform transformation of austenite
to martensite—quench thoroughly and uniformly. Do not
short cycle the tempering operation. Double and triple tem-
per air-hardening steels. Employ subzero and cryogenic
quenching to get complete transformation (austenite to mar-
tensite).
Growth in at least one dimension is expected in proper heat
treatment. Refer to Table 8-9 for the expected growth of various
types of tool steels in heat treatment. Note that shrinkage may
develop on a given dimension if full transformation of austenite
to martensite is not realized. It can also occur on a dimension
after proper heat treatment as a result of part geometry.

189

Ch08.pmd 189 10/29/02, 11:25 AM


Tool and Die Making Troubleshooter

Problem: Cracking During Hardening


Possible causes and remedies:
• Inaccurate furnace controls allow over- or under-austen-
itization—check and calibrate furnace controls.
• Design configurations create stress raisers—simplify part
designs that incorporate stress raisers.
• Severe mechanical stress concentrations—stress relieve parts
that have been severely cold worked before heat treatment.
• Carburization and decarburization—anneal and/or harden
in furnaces with controlled neutral atmospheres.
• Grain coarsening from overheating and oversoaking make
steel brittle—preheat thoroughly before hardening. Do not
oversoak or undersoak.
• Nonuniform quenching—quench uniformly. Use air-harden-
ing steels whenever possible.
• Short cycling (insufficient) tempering after quenching—tem-
per immediately after quenching. Avoid tempering at tem-
perature ranges under 400° F (204° C). Double and triple
temper air-hardening steels.

REFERENCES
300° F Below, Inc. Undated. “Deep Cryogenic Tempering,” bro-
chure. Decatur, IL: 300° F Below, Inc.
Bethlehem Steel Corp. 1936. Alloy and Special Steels. Catalog
#107. Bethlehem, PA: Bethlehem Steel Corp., pp. 193–195.
——. 1950. “A Heat Treater’s Lament.” Tool Steel Topics, No. 29,
July. Bethlehem, PA: Bethlehem Steel Corp.
——. 1954. Bethlehem Tool Steels. Bethlehem, PA: Bethlehem Steel
Corp., D3 pp. 68–71, H21 pp. 139–143, S2 pp. 88–92.
——. 1965. “Distortion of Tool Steels in Heat Treatment.” Book-
let 2154. Bethlehem, PA: Bethlehem Steel Corp., pp. 2–8.
——. 1967. “An Evaluation of Type 321 Stainless-steel Foil Wrap
in Hardening Tool-steel Parts.” Folder 2384. Bethlehem, PA:
Bethlehem Steel Corp.

190

Ch08.pmd 190 10/29/02, 11:25 AM


Chapter 8: Heat Treatment

——.1976. “Bearcat AISI S7 Bethlehem Shock-resisting Tool


Steel.” Folder 2412-B, October. Bethlehem, PA: Bethlehem Steel
Corp.
——. 1977. “BTR AISI Type O1 Bethlehem Cold-work Tool Steel.”
Folder 2321-B. Bethlehem, PA: Bethlehem Steel Corp.
——. 1978a. “Carbon-vanadium AISI Type W2 Bethlehem Water-
hardening Tool Steel Data Sheet.” Folder 2722-A. Bethlehem, PA:
Bethlehem Steel Corp.
——. 1978b. “Omega AISI Type S5 Bethlehem Shock-resisting
Tool Steel.” Folder 2151-E. Bethlehem, PA: Bethlehem Steel Corp.
——. 1978c. “A-6 AISI Type A6 Bethlehem Cold-work Tool Steel.”
Folder 3718. Bethlehem, PA: Bethlehem Steel Corp.
——. 1978d. “Lehigh H AISI Type D2 Bethlehem Cold-work Tool
Steel.” Folder 2323-B. Bethlehem, PA: Bethlehem Steel Corp.
——. 1978e. “Cromo-high V AISI Type H13 Bethlehem Hot-work
Tool Steel.” Folder 2113-C. Bethlehem, PA: Bethlehem Steel Corp.
——. 1978f. “M2 High-speed AISI Type M2 High-speed Steel.”
Folder 2484-B. Bethlehem, PA: Bethlehem Steel Corp.
——. 1978g. “Bethadur 420 AISI Type 420 Bethlehem Plastic
Molding Tool Steel.” Folder 3043-A. Bethlehem, PA: Bethlehem
Steel Corp.
——. 1980. Modern Steels and Their Properties. Handbook 3310.
Bethlehem, PA: Bethlehem Steel Corp., pp. 89–161.
——. 1981a. “A-H5 AISI Type A2 Bethlehem Cold-work Tool
Steel.” Folder 2322-D. Bethlehem, PA: Bethlehem Steel Corp.
——. 1981b. The Tool Steel Troubleshooter. Handbook 2828-C.
Bethlehem, PA: Bethlehem Steel Corp., pp. 62–63.
Haga, L. J. 1974. Atmosphere, Quench, and Draw Treating. Grand
Rapids, MI: L. J. Haga.
——. 1989a. “Process Control: Surface Protection and Endother-
mic Atmospheres.” Heat Treating, February. New York: Fairchild
Publications.

191

Ch08.pmd 191 10/29/02, 11:25 AM


Tool and Die Making Troubleshooter

——. 1989b. “Surface Protection, Continued: Exo, Salts, and Other


Methods.” Heat Treating, March. New York: Fairchild Publica-
tions.
——. Undated a. “A-HT Bethlehem Cold-work Tool Steel.” Folder
3002. Bethlehem, PA: Bethlehem Steel Corp.
——. Undated b. “O-6 Graphitic AISI Type O6 Bethlehem Cold-
work Tool Steel.” Folder 2336. Bethlehem, PA: Bethlehem Steel
Corp.
——. Undated c. “67 Chisel AISI S1 Bethlehem Shock-resisting
Tool Steel.” Bethlehem, PA: Bethlehem Steel Corp.
——. Undated d. “How to Heat Treat Large Sections of Air-hard-
ening Tool Steels.” Brochure #2374-A. Bethlehem, PA: Bethlehem
Steel Corp.
Kimble, William H. 1977. “Vacuum ...Is It Really Nothing?”
Cranston, RI: C.I. Hayes, Inc.
N. E. Slavin and Co. 1972. “Tool Steel Decarburization.” Heat
Treating, April. New York: Fairchild Publications.
Sullivan, J. W. 1974. “Vacuum Keeps Tools from Carburizing, De-
carburizing.” Metal Progress, September (now Advanced Materi-
als and Processes). Materials Park, OH: ASM International, pp.
109–110.
Sweeney, Thomas P., Jr. 1986. “Deep Cryogenics: The Great Cold
Debate.” Heat Treating, February. New York: Fairchild Publica-
tions.
Teledyne Sterling Hunt. Undated. “Dual-Foil Brand Heat-treat-
ing Wraps,” brochure. South Dartmouth, ME: Teledyne Sterling
Hunt.
Teledyne Vasco. 1968. High-speed Steels. Latrobe, PA: Teledyne
Vasco.

192

Ch08.pmd 192 10/29/02, 11:25 AM


Chapter 9: Hardness Testing

9
Hardness Testing

One of the most important aspects of tool-and-die construction


is determining material hardness. This chapter describes proce-
dures and techniques for three of the most popular hardness test-
ing methods: file, Rockwell C, and Brinell.

FILE TESTING
It has long been common practice for die makers and heat
treaters to use hardened files to check the effectiveness of heat
treatment. The test is performed by lightly pressing a file of
standardized or known hardness against the part and pushing it
with a forward motion over the surface (Figure 9-1).
An experienced file tester can feel the “bite” or “slip” of the file
through his fingers as pressure is applied with the file against the
workpiece. If the file does not bite, the part is considered to be
“file hard” because it is as hard or harder than the file being used

Figure 9-1. Recommended motion directions for file testing.

193

Ch09.pmd 193 10/29/02, 11:26 AM


Tool and Die Making Troubleshooter

for the test. Since most tooling is hardened in the range of 40–65 on
the Rockwell C (HRC) scale, testing becomes a simple matter of
subjecting the part to a series of files that vary in hardness (that is,
going from a high-hardness-level file that will bite to a soft file that
will not bite into the steel). File sets are available that are hard-
ened and tempered in five-point increments from 40–65 HRC.
Here are some tips for performing file hardness tests:
• Be sure the base-metal hardness is being checked and not an
area that has been decarburized. Grind the surface to remove
decarb, rust, or scale.
• The safest procedure is to use special test files made by file
manufacturers.
• Speed, pressure, and angle of contact affect the results of the
test. Use a slow, forward motion with gentle pressure. Slow
speed, firm pressure, and a constant angle of contact give the
best results.
• A polished surface is sometimes difficult to bite into, while a
rough or ground surface on the same part may appear to bite
easily.
• Starting with a file known to be harder than the material
being checked, run the file over the workpiece. If it marks
the workpiece, the file is harder. Move down in file hardness.
Continue checking with progressively softer files until the
workpiece cannot be marked. The approximate hardness of
the workpiece is between the file that marks the workpiece
and the file that slips without marking.
• Generally, steels that are easy to file (annealed or soft) are 20–
40 HRC. Those difficult to file (but let the file bite) are 45–55
HRC. Steels that do not let the file bite are 60 HRC and higher.
• Worn files lose their reliability. A general rule is to discard
them after the depth of their cutting teeth has been reduced
by 50% or more.
The file test can be a valuable check for decarburization, soft
spots, and hardness comparisons. It also can be used for deter-
mining hardness levels in tools that are either too large to be tested
on a bench tester or that are otherwise inaccessible.

194

Ch09.pmd 194 10/29/02, 11:26 AM


Chapter 9: Hardness Testing

ROCKWELL TESTING
The Rockwell C scale test is the most common of the hardness
tests used on tooling that has been heat treated to a hardness over
40 HRC. A 330-lbf (150-kgf) load is squeezed into the workpiece
by the testing machine. The resistance to penetration is measured
by a spheroconical diamond penetrator as it is pressed into the
workpiece by the weight load (Figure 9-2). The increment of in-
dentation depth produced by the weight load is the basis for the
Rockwell hardness number.
Some tips on Rockwell testing procedures include:
• Check for a 330-lbf (150-kgf) weight load.
• Be sure the anvil is properly seated and clean.
• Check the diamond; be sure it is not tightened too hard in its
sleeve.

Figure 9-2. Closeup of the Rockwell penetrator (Instron/Wilson/Shore In-


struments 1988).

195

Ch09.pmd 195 10/29/02, 11:26 AM


Tool and Die Making Troubleshooter

• Check the accuracy of the machine with a certified testing


block.
• Never make hardness indentations on both sides of a test
block.
• Check the workpiece. The surface being tested should be clean
and smooth. A parallel flat surface should be presented to
the penetrator. Remove burrs, coatings, and rough machine
marks.
• Always take at least four hardness tests in a particular area.
Omit the first one because the penetrator may be seating
itself. If three readings are within ±1 point HRC, the results
can be averaged and reported. If not, additional readings
should be taken.
• Never retest in the same indentation because it will be work
hardened.
• Hardness impressions should not be too close to one another.
The distance between indentations should be at least three
times the diameter of an indentation.
• Avoid taking readings too close to edges and grooves and holes.
Stay at least two and a half times the diameter of the hard-
ness indentation away from these areas (Instron/Wilson/Shore
Instruments 1988).

BRINELL TESTING
The Brinell hardness test is an indentation test using calibrated
machines to force a hard ball, under specified conditions, into the
surface of a material. After removal of the test force, the diameter
of the impression is determined with a measuring microscope and
converted into a Brinell hardness number. The measuring micro-
scope should be graduated to read at least in increments of 0.0002
in. (0.005 mm).
Generally, material in the annealed condition or low hardness
range of 375 HB (hardness Brinell) or less is more accurately
checked for hardness with a Brinell test. Material over 375 HB
(40 HRC) should be tested for hardness with a Rockwell C test.
Figure 9-3 illustrates the four steps in Brinell hardness testing.

196

Ch09.pmd 196 10/29/02, 11:26 AM


Chapter 9: Hardness Testing

Figure 9-3. Four steps in Brinell hardness testing.

Table 9-1 is a hardness-conversion table that provides Brinell


hardness numbers corresponding to various diameters of impres-
sion for 6,600 lbf (3,000 kgf) loads with a 0.39-in. (10-mm) diam-
eter ball used for making the hardness impression. Other loads
may be used in Brinell hardness testing, such as 1,100 and 3,300
lbf (500 and 1,500 kgf). The choice of these lighter loads depends
on the hardness and thickness of the material being tested.
Charts for converting the mean diameter of the hardness im-
pression to a Brinell hardness number when a 0.39-in. (10-mm)
diameter ball is used may be found in either ASTM Specification
E-10, “Standard Test Method for Brinell Hardness of Metallic
Materials,” or ASTM Specification A-370, “Standard Test Meth-
ods and Definitions for Mechanical Testing of Steel Products.”
For material under 450 HB, a 0.39-in. (10-mm) steel ball is used.
For harder materials, a 0.39-in. (10-mm) tungsten-carbide ball is
used. As a general rule, the test specimen thickness should be at
least 10 times the depth of the indentation. If a bulge appears on
the underside of the test piece after a hardness impression is made,
that test result should be discarded.

197

Ch09.pmd 197 10/29/02, 11:26 AM


Ch09.pmd

Tool and Die Making Troubleshooter


Table 9-1. Hardness conversion
Brinell Indent Approximate
Diameter, in. Brinell Rockwell Rockwell Rockwell Superficial Vickers Tensile Strength,
(mm) Number* A** B† C†† 30 N Number ksi§ (kgf/cm2)
— — 86.5 — 70.0 86.0 1076 —
— — 86.0 — 69.0 85.0 1004 —
— — 85.6 — 68.0 84.4 940 —
198

— — 85.0 — 67.0 83.6 900 —


— 757 84.4 — 65.9 82.7 860 —
0.089 (2.25) 745 84.1 — 65.3 82.2 840 —
— 722 83.4 — 64.0 81.1 800 —
— 710 83.0 — 63.3 80.4 780 —
0.093 (2.35) 682 82.2 — 61.7 79.0 737 —
198

0.094 (2.40) 653 81.2 — 60.0 77.5 697 —


0.097 (2.45) 627 80.5 — 58.7 76.3 667 323 (22,709)
0.098 (2.50) 601 79.8 — 57.3 75.1 640 309 (21,725)
0.100 (2.55) 578 79.1 — 56.0 73.9 615 297 (20,881)
0.102 (2.60) 555 78.4 — 54.7 72.7 591 285 (20,038)
0.104 (2.65) 534 77.8 — 53.5 71.6 569 274 (19,264)
0.106 (2.70) 514 76.9 — 52.1 70.3 547 263 (18,491)
10/29/02, 11:26 AM

0.108 (2.75) 495 76.3 — 51.0 69.4 528 253 (17,788)


0.110 (2.80) 477 75.6 — 49.6 68.2 508 243 (17,085)
0.112 (2.85) 461 74.9 — 48.5 67.2 491 235 (16,522)
0.114 (2.90) 444 74.2 — 47.1 65.8 472 225 (15,819)
0.116 (2.95) 429 73.2 — 45.7 64.6 455 217 (15,257)
0.118 (3.00) 415 72.8 — 44.5 63.5 440 210 (14,765)
0.120 (3.05) 401 72.0 — 43.1 62.3 425 202 (14,202)
0.122 (3.10) 388 71.4 — 41.8 61.1 410 195 (13,710)
Ch09.pmd

Table 9-1. (continued)


Brinell Indent Approximate
Diameter, in. Brinell Rockwell Rockwell Rockwell Superficial Vickers Tensile Strength,
(mm) Number* A** B† C†† 30 N Number ksi§ (kgf/cm2)
0.124 (3.15) 375 70.6 — 40.4 59.9 396 188 (13,218)
0.126 (3.20) 363 70.0 — 39.1 58.7 383 182 (12,796)
0.128 (3.25) 352 69.3 110.0 37.9 57.6 372 176 (12,374)
199

0.130 (3.30) 341 68.7 109.0 36.6 56.4 360 170 (11,952)
0.132 (3.35) 331 68.1 108.5 35.5 55.4 350 166 (11,671)
0.134 (3.40) 321 67.5 108.0 34.3 54.3 339 160 (11,249)
0.136 (3.45) 311 66.9 107.5 33.1 53.3 328 155 (10,898)
0.138 (3.50) 302 66.3 107.0 32.1 52.2 319 150 (10,546)
0.140 (3.55) 293 65.7 106.0 30.9 51.2 309 145 (10,195)
199

0.142 (3.60) 285 65.3 105.5 29.9 50.3 301 141 (9,913)
0.144 (3.65) 277 64.6 104.5 28.8 49.3 292 137 (9,632)
0.146 (3.70) 269 64.1 104.0 27.6 48.3 284 133 (9,351)
0.148 (3.75) 262 63.6 103.0 26.6 47.3 276 129 (9,070)
0.150 (3.80) 255 63.0 102.0 25.4 46.2 269 126 (8,859)
0.152 (3.85) 248 62.5 101.0 24.2 45.1 261 122 (8,578)

Chapter 9: Hardness Testing


0.154 (3.90) 241 61.8 100.0 22.8 43.9 253 118 (8,296)
10/29/02, 11:26 AM

0.156 (3.95) 235 61.4 99.0 21.7 42.9 247 115 (8,085)
0.157 (4.00) 229 60.8 98.2 20.5 41.9 241 111 (7,804)
0.159 (4.05) 223 59.7 97.3 18.8 — 234 —
0.161 (4.10) 217 59.2 96.4 17.5 — 228 105 (7,382)
0.163 (4.15) 212 58.5 95.5 16.0 — 222 102 (7,171)
0.165 (4.20) 207 57.8 94.6 15.2 — 218 100 (7,031)
0.167 (4.25) 201 57.4 93.8 13.8 — 212 98 (6,890)
Ch09.pmd

Tool and Die Making Troubleshooter


Table 9-1. (continued)
Brinell Indent Approximate
Diameter, in. Brinell Rockwell Rockwell Rockwell Superficial Vickers Tensile Strength,
(mm) Number* A** B† C†† 30 N Number ksi§ (kgf/cm2)
0.169 (4.30) 197 56.9 92.8 12.7 — 207 95 (6,679)
0.171 (4.35) 192 56.5 91.9 11.5 — 202 93 (6,539)
0.173 (4.40) 187 55.9 90.7 10.0 — 196 90 (6,328)
200

0.175 (4.45) 183 55.5 90.0 9.0 — 192 89 (6,257)


0.177 (4.50) 179 55.0 89.0 8.0 — 188 87 (6,117)
0.179 (4.55) 174 53.9 87.8 6.4 — 182 85 (5,976)
0.181 (4.60) 170 53.4 86.8 5.4 — 178 83 (5,836)
0.183 (4.65) 167 53.0 86.0 4.4 — 175 81 (5,695)
0.185 (4.70) 163 52.5 85.0 3.3 — 171 79 (5,554)
200

0.189 (4.80) 156 51.0 82.9 0.9 — 163 76 (5,343)


0.193 (4.90) 149 49.9 80.8 — — 156 73 (5,132)
0.197 (5.00) 143 48.9 78.7 — — 150 71 (4,992)
0.201 (5.10) 137 47.4 76.4 — — 143 67 (4,711)
0.205 (5.20) 131 46.0 74.0 — — 137 65 (4,570)
0.209 (5.30) 126 45.0 72.0 — — 132 63 (4,429)
0.213 (5.40) 121 43.9 69.8 — — 127 60 (4,218)
0.217 (5.50) 116 42.8 67.6 — — 122 58 (4,078)
10/29/02, 11:26 AM

0.220 (5.60) 111 41.9 65.7 — — 117 56 (3,937)


Values in italics are beyond the normal range and given for information only.
* 0.39-in. (10-mm) diameter, tungsten-carbide ball forced into the material at 6,614 psi (3,000 kgf/cm2)

** 132 lb (60 kg) brale 220 lb (100 kg), 1/16 in. (1.6 mm) ball
†† §
331 lb (150 kg) brale 1,000 psi or 1,000 lbf/in.2
(The Timken Company 1999)
Chapter 9: Hardness Testing

REFERENCES
Brandt, Daniel A. 1985. Metallurgy Fundamentals. South Hol-
land, IL: The Goodheart-Wilcox Co., Inc., pp. 84–85.
Instron/Wilson/Shore Instruments. 1988. Fundamentals of
Rockwell Hardness Testing. Canton, MA: Instron/Wilson/Shore
Instruments, p. 1.
The Timken Company. 1999. Practical Data for Metallurgists.
Canton, OH: The Timken Company, pp. 116–117.

201

Ch09.pmd 201 10/29/02, 11:26 AM


Ch09.pmd 202 10/29/02, 11:26 AM
Chapter 10: Wear Enhancement Treatments

10
Wear Enhancement Treatments

This chapter first traces the transition of tooling steels from


the plain carbon variety to highly alloyed steels, including high-
speed steels, as part of the quest for better, longer-running tools.
It then explains the evolution and use of surface coatings and treat-
ments for further extending the life of tools by improving their
wear and corrosion resistance, and lubricity (less likely to seize
and gall).

SURFACE TREATMENTS ARE NOT COVERUPS


The surface coatings and treatments described here are designed
strictly for tooling-wear enhancement. They are not a cure-all for
tooling that has been poorly designed, poorly heat treated, or not
performed well due to improper material selection or construc-
tion practices.
Coatings require a firm, solid substrate for support. Also, tool-
and-die steels used as substrate materials must be carefully se-
lected. For example, low-carbon shock-resisting steel should be
used wherever impact and toughness preclude the use of more
wear-resistant steels like D2.
When the proper type of surface coating or treatment is care-
fully selected, the results can be extremely rewarding and profit-
able. Some benefits are:
• longer uninterrupted production runs,
• reduced tool maintenance,

203

Ch10.pmd 203 10/29/02, 11:26 AM


Tool and Die Making Troubleshooter

• reduced use of manufacturing lubricants,


• increased tool life and performance, and
• improved part quality.
For maximum benefit, surface coatings and treatments (such as
chrome plating, ion nitriding, thermal diffusion, etc.) should be used
by toolmakers to preserve a good working tool at its optimal state,
as well as reduce die wear, reduce the need for lubrication, and
protect the working surfaces of the tools from unnecessary grind-
ing and stoning.
Surface treatments should never be used as a remedy for a poorly
produced tool. Doing so will merely preserve an unacceptable con-
dition and may even intensify it. Furthermore, surface treatments
should be used only on tooling materials in optimal condition for
an application.

A FIRM FOUNDATION
The properties of the substrate or base metal are extremely
important for good tool life. To support the treated or coated sur-
face, the key qualifications are:
• Good tool design, that is, designs without built-in stress rais-
ers that can cause the tool to crack in heat treatment or fail
early in service.
• Proper selection of tool steel and working hardness level to
match the necessary physical and metallurgical properties
for the application and thus maximize performance.
• Proper attention to and implementation of good heat-treat
practices to achieve optimal physical and metallurgical prop-
erties.
Many tools and dies fail early in service because of built-in de-
sign faults and/or inadequate steel selection for the application.
However, the most likely cause of failure is improper or inadequate
heat treatment. Specifically, the key causes are failure to control
the surface chemistry of the steel, quenching too cold or ineffec-
tive tempering, and quenching from excessively high austenitizing
temperatures. To perform well, tools and dies must be properly
heat treated, whether or not the tool’s wear surface is enhanced.

204

Ch10.pmd 204 10/29/02, 11:26 AM


Chapter 10: Wear Enhancement Treatments

Tools and dies should be carefully produced and proven by try-


out to work properly in production before being preserved by some
type of surface-wear enhancement—an approach strongly recom-
mended by most vendors of surface-coating or treatment processes.
Many times tools require minor changes during tryout to allow
them to perform properly. All such changes should be made be-
fore surface enhancement. This recommendation is made even
for surface-enhancement processes such as thermal diffusion.
During this treatment, the process temperature is so high—parts
are immersed in molten salt at 1,600–1,900° F (871–1,038° C)—
that the treated tool ends up being retreated during the quench-
ing that follows completion of the soaking time in the molten salt.
Thus, toolmakers are advised to consult their surface-treatment
vendor for suggestions and/or recommendations for proper tool
condition prior to surface treatment.

PROGRESS IN WEAR ENHANCEMENT


From the earliest days of mass manufacturing, users of tools
and dies looked for ways to increase the length of production
runs and avoid shutting down for tooling changes. The incen-
tives were obvious: reduce lost production time and/or high main-
tenance costs.
For centuries, carbon-steel (approximately 1% carbon) tooling
was heat treated to produce martensite and then tempered to re-
duce brittleness. Then, in the late 19th century, steel producers
began to add other elements (such as chromium) to improve steel
properties—primarily wear resistance. These additions also al-
lowed the steel to be more safely hardened, with less distortion by
quenching in oil rather than in water. The limitation of these
materials was primarily their loss of cutting hardness at tempera-
tures exceeding the relatively low tempering temperatures used
in their heat treatment.
Around 1900, high-speed steels were developed. These steels—
heat treated by quenching from near their melting temperatures
and tempered above 1,000° F (538° C)—could be heated by the
friction on the cutting tool to a temperature of almost red heat
without losing cutting hardness. One famous metallurgist in the

205

Ch10.pmd 205 10/29/02, 11:26 AM


Tool and Die Making Troubleshooter

early 1900s joked that high-speed steel “was the only substance
then known that could get hot without losing its temper.”
Further developments led to other wear-resisting and shock-
resisting steels. The increased alloy content of many of these steels
allowed them to be heat treated in the safest manner possible,
that is, by air hardening.
As production processes were coincidentally developed, ways
were sought to minimize the downtime resulting from breakage,
premature wear, and the need for excessive die maintenance.
Materials were selected for improved wear, as well as resistance
to seizing, galling, and corrosion. Lubricants and other compounds
were added to the process to get the most out of the tooling. How-
ever, something more than simply changing from one grade of
steel to another was needed to further increase wear and other
desirable service characteristics. Thus, the search for wear-en-
hancement coatings and treatments began.

Early Wear Treatments


In addition to early efforts to extend tooling life with alloyed
and special-purpose tool steels, some researchers also attempted
to improve service life by altering tooling surfaces. In the early to
mid-1900s, tooling surfaces were treated by such methods as car-
burizing, gas nitriding, cyaniding, carbonitriding, flame harden-
ing, and chrome plating. These processes improved wear by either
increasing the surface hardness, changing the surface composi-
tion, or coating it with another material.

Carburizing
Carburizing raises the steel surface’s carbon content to a value
higher than its base metal or substrate by subjecting the workpiece
to a carbon-rich atmosphere at a temperature of 1,700° F (927°
C). Upon quenching, the resulting surface is harder and more wear
resistant.

Gas Nitriding
In gas nitriding, the part is heated to a temperature of 900–
1,150° F (482–621° C) in an atmosphere of ammonia gas and dis-

206

Ch10.pmd 206 10/29/02, 11:26 AM


Chapter 10: Wear Enhancement Treatments

sociated ammonia for an extended period of time. This imparts a


thin, very hard case from the formation of nitrides. Nitrided parts
have exceptional wear resistance with little tendency to gall and
seize, making them particularly serviceable in applications involv-
ing metal-to-metal wear. They also have high resistance to fatigue.

Cyaniding
Cyaniding is performed in a bath of sodium cyanide heated to a
temperature slightly above the transformation range, 1,350–1,600°
F (732–871° C) depending on the grade. Because of the presence
of nitrides, this results in a thin case of high hardness that has
superior wear resistance, approaching that of a nitrided case.

Carbonitriding
Carbonitriding, also known as gas cyaniding, is similar to
cyaniding except that the absorption of carbon and nitrogen is
accomplished by heating to 1,200–1,650° F (649–899° C) in a gas-
eous atmosphere containing hydrocarbons and ammonia. The
carbonitrided parts are then quenched and tempered, resulting
in a hard, wear-resistant case.

Flame Hardening
Flame hardening involves rapid heating with a direct, high-
temperature gas flame that heats the surface layer of the part
above the transformation range. This is followed by cooling at a
rate to accomplish the desired hardening. The process does not
alter surface chemistry, but does result in a higher hardness than
that present in the substrate. The flame-hardened surface dis-
plays improved wear resistance simply because of its higher hard-
ness (Bethlehem Steel Corp. 1980).

Chrome Plating
Chrome plating is electrolytically deposited on a die surface at
a temperature of approximately 140° F (60° C). The plating seals
the surface, and with proper maintenance, eliminates wear. The
lubricity of the chrome coating increases metal flow in drawing

207

Ch10.pmd 207 10/29/02, 11:26 AM


Tool and Die Making Troubleshooter

operations and can significantly reduce the need for drawing com-
pounds. The depth of coating is usually 0.0005 in. (0.013 mm) on
flat areas and 0.0010–0.0015 in. (0.025–0.038 mm) on radii. As-
plated hardness is approximately 62–64 HRC. Several proprietary
vendors of chrome plating claim that their special processing re-
sults in hardness at the top end of this range.

Advancements
In the latter half of the 20th century, a host of processes were
developed to improve the working surfaces of tools and dies. These
include physical-vapor deposition (PVD), chemical-vapor deposi-
tion (CVD), thermal diffusion, ferritic nitrocarburizing, ion im-
plantation, ion nitriding, deep cryogenic tempering (DCT),
plasma-assisted CVD, plasma-source ion implantation (PSII), and
micro-plasma deposition.

Physical-vapor Deposition (PVD)


PVD is a low-temperature, hard-coating process carried out
under vacuum. The workpieces to be coated are placed in a reac-
tor evacuated to vacuum conditions. TiN-PVD coatings are pro-
duced from a glow-discharge plasma containing ions of titanium
and nitrogen. The titanium ions are produced by various activated-
reactive-evaporation and reactive-sputtering processes. The en-
ergy required to produce the desired chemical reaction is supplied
by a high-energy electric field.
The process is line-of-sight: surfaces to be coated must be placed
in the path of the coating source. Coatings can be applied to sur-
faces preferentially by positioning tools in the reactor, and mask-
ing can be used to keep certain surfaces coating-free.
PVD coatings are titanium nitride (TiN), titanium carbonitride
(TiCN), chromium nitride (CrN), and chromium carbide (CrC).
Coating thicknesses range from 0.00008–0.00020 in. (2–5 ␮m).
Because the process is carried out at a low temperature (that
is, below the normal tempering temperature of most tool and die
steels), no postcoating heat treatment is required. Tools with tight
tolerances are better coated via PVD than CVD. The PVD process
is the preferred method for coating cutting tools.

208

Ch10.pmd 208 10/29/02, 11:26 AM


Chapter 10: Wear Enhancement Treatments

Chemical Vapor Deposition (CVD)


CVD is presently the most widely used process by tool manu-
facturers for improving the surface of tooling. It is performed by
placing the workpieces to be coated inside a reactor heated to
1,850–2,000° F (1,010–1,093° C). Reactant gases, which carry the
components of the coating, are fed into the reactor under con-
trolled conditions. Coating thicknesses are very thin and usually
range from 0.00024–0.00040 in. (6–10 µm).
Coatings consist of thin films of such materials as TiN, TiC,
TiCN, Al2O3, sometimes in layered combinations. CVD is the pre-
ferred method for coating tools subjected to severe compression
stresses, such as those used in extrusion, forming, flanging, and
drawing.
Because the process is conducted at high temperature, CVD-
coated tools must be subsequently heat treated to produce the
required substrate hardness. Hardened tools that are subjected
to the coating process must be heat treated again to restore hard-
ness and metallurgical structure (Richter Precision, Inc. 1993;
Clavel Undated).

Thermal Diffusion
Thermal diffusion is a high-temperature, surface-modification
process that forms a carbide layer on carbon-containing materi-
als (0.3% minimum carbon), such as steels, nickel alloys, cobalt
alloys, and cemented carbides. The diffused layer measures from
0.0001–0.0008 in. (2.5–20 ␮m). Thermal-diffusion-processed ma-
terials exhibit high hardness and excellent resistance to wear, sei-
zure, and corrosion.
The process is performed by immersing parts in a molten salt
bath at temperatures of 1,660–1,900° F (904–1,038° C). Carbide
constituents, dispersed in the salt, combine with the carbon at-
oms contained in the tooling substrate to form a carbide layer.
The process produces layers of vanadium, niobium, and chromium
carbide, depending on the carbide-forming elements used in the
salt bath. Quenching in air and tempering follow the removal of
treated parts from the salt bath. Steels whose austenitizing tem-
perature exceed the maximum temperature of 1,900° F (1,038° C)

209

Ch10.pmd 209 10/29/02, 11:26 AM


Tool and Die Making Troubleshooter

are frequently used for this process. Heat treatment in a vacuum


or salt bath after thermal-diffusion processing is necessary (Smith
1990; Arter 1990).

Ferritic Nitrocarburizing
Ferritic nitrocarburizing is a deep-hardening, surface-treating
process conducted in either atmosphere or fluidized-bed furnaces.
The hardened layer on a workpiece is produced by a mixture of
nitrogen, ammonia, and hydrocarbon gases introduced into the fur-
nace. Carbon is combined with nitrogen on the surface of the work-
piece to produce a very wear-resistant 70+ HRC coating. The
process is a low-temperature one, with operating temperatures
ranging from 600–1,200° F (316–649° C). The total diffusion zone
depth of treated surfaces is 0.003–0.005 in. (0.08–0.13 mm) for a
normal cycle, 0.005–0.010 in. (0.13 to 0.25 mm) for a double cycle,
and 0.010–0.015 in. (0.25–0.38 mm) for a triple cycle. Dimensional
change for a normal cycle is 0.0001–0.0002 in. (2.5–5 ␮m) per side
(Dynamic Metal Treating, Inc. Undated a, b).

Deep Cryogenic Tempering (DCT)


DCT subjects tools and dies to deep-freezing temperatures of
–300° F (–184° C), the temperature of liquid nitrogen. Normally,
deep freezing of tools and dies is performed at temperatures of
–120° F (–84° C).
DCT equipment achieves a controlled dry thermal treatment.
“Controlled” simply means that the process is performed accord-
ing to a precise, prescribed timetable. The descent, soak, and as-
cent modes of the operating cycle are computer controlled.
The major benefit of the DCT process is improved wear resis-
tance. It affects not only the surface of the tool or die, but the
entire part. Therefore, the benefit of the process is not lost through
wear and regrinds. This is reportedly accomplished through the
complete transformation of all soft retained austenite to hard
martensite throughout the substrate, and through the precipita-
tion of fine carbide, which enhances the strength and toughness
of the martensitic matrix. The process is said to provide a denser
structure, resulting in a larger surface area that reduces friction,
heat, and wear. Other claims include increased tensile strength,

210

Ch10.pmd 210 10/29/02, 11:26 AM


Chapter 10: Wear Enhancement Treatments

toughness, and stability. The process reportedly improves the prop-


erties of all heat-treated tool steels except for the water-harden-
ing grades. Success has been claimed on some sintered carbide
grades, cast steel, and cast iron (Paulson 1993).

Ion Implantation
Ion implantation is a process of introducing atoms of alloying
elements into the surface layer of a metallic workpiece. The at-
oms are accelerated to high energies that penetrate the surface of
the metal to depths ranging from 0.400–39.400 ␮in. (0.01–1.00
␮m), depending on the energy of the atom. The process is used to
create a thin alloy layer on the workpiece surface.
Ion implantation is conducted under vacuum and at low tem-
perature. The only heating that occurs is due to the energetic at-
oms colliding with the atoms of the base material. The maximum
temperature of the workpiece seldom exceeds 400° F (204° C), and
can be lowered by controlling the rate of implantation.
In operation, atoms of the desired alloying elements are fed
into an ion-source assembly where they are ionized by an electri-
cal discharge. If the element is in a gaseous state, such as nitro-
gen, the purified gas is fed directly into the ion source. If the
element is a solid, such as chromium, it is first vaporized and then
ionized. High voltage is used to accelerate the ions from the source.
Ion implantation has been applied most heavily to the treat-
ment of forming tools, but is used on many types of tooling. By
far, the element most implanted is nitrogen. Nitrogen-ion implan-
tation results in an abrasion-resistant, hard-case layer with an
effective hardness of 80–90 HRC which can withstand surface tem-
peratures up to 600° F (316° C).
The process causes no problems with thermal distortion, melt-
ing, or residual heat-treatment effects. Its limitation is that it is a
line-of-sight process, making it unsuitable for some complex ge-
ometries (Deutchman and Partyka 1993; Beam Alloy Corp. 1989).

Plasma-source Ion Implantation (PSII)


PSII is a newer, low-temperature technology for producing a
hard, wear-resisting layer over the entire surface of irregularly
shaped components. The part to be treated by PSII is placed in a

211

Ch10.pmd 211 10/29/02, 11:26 AM


Tool and Die Making Troubleshooter

vacuum chamber in which there is a plasma containing the proper


ions. The part is then pulse-biased to a high negative voltage, caus-
ing the positively charged ions to accelerate at high velocity to-
ward its surface. The subsequent implantation of ions produces
chemical and microstructural changes in the steel, leading to rel-
evant surface modification of the part.
PSII is not a coating, but rather a combination of impregnation
into the applied base metal, plus an integrated layer on the sur-
face. It virtually becomes one with the base metal. Similarly, PSII
is not one treatment but a variety of treatments, all applied at
room temperature.
The following PSII options are currently available:
• A diamond-like treatment is impregnated into the base metal
up to 3.90 ␮in. (0.1 ␮m) thick and deposited on the surface
up to 278 ␮in. (7.6 ␮m). No microcracks and tremendous lu-
bricity are the result. Hardness is equivalent to 88 HRC up
to 3.90 ␮in. (0.1 ␮m).
• Titanium mixed with boron is impregnated into the base
metal and deposited on the surface up to 27.60 ␮in. (0.7 ␮m)
into the base metal, yielding 50% higher hardness than tita-
nium nitride, with superior lubricity and bonding. Hardness
is equivalent to 98 HRC.
• After hard-chrome plating, nitrogen is implanted 3.90 ␮in.
(0.1 ␮m) into the plated layer, producing extremely high hard-
ness (equivalent to 96 HRC).
• Implanted chrome treatment may be done at room tempera-
ture to steel, forming a stainless-like surface with high hard-
ness equivalent to 72 HRC up to 3.90 ␮in. (0.1 ␮m) into the
base metal.
All these treatments result in high surface hardness and lubricity
with no microcracks and extremely good corrosion protection
(Reeber and Sridharan 1994).

Ion Nitriding
Ion nitriding is a method of surface hardening for producing
nitrided cases. It uses glow-discharge technology to generate ni-
trogen ions to the surface of a metallic part for diffusion. It is a

212

Ch10.pmd 212 10/29/02, 11:26 AM


Chapter 10: Wear Enhancement Treatments

thermal/chemical process, accomplished in a vacuum furnace


where high electrical energy is used to form a plasma through
which molecules dissociated by discharge are accelerated toward
the surface of the workpiece. These molecules clean the surface
and provide active nitrogen for diffusion. The process penetrates
from 0.005–0.020 in. (0.13–0.50 mm), and surface hardness is in-
creased to 60–65 HRC.
The process results in improved resistance to wear, fatigue, and
corrosion, as well as excellent antiscoring and antigalling proper-
ties. Better lubricity is achieved through a small increase in sub-
microscopic roughness, which helps retain lubricants. The lower
processing temperature—700–1,200° F (371–649° C)—results in
less distortion than conventional nitriding.

Micro-plasma Deposition
Micro-plasma deposition, also called plasma thermal spraying,
is a process for producing coatings of a wide variety of composi-
tions, the most popular of which are tungsten carbide and chro-
mium carbide. In the process, the desired deposition elements, in
powder form, are fed into a high-velocity plasma stream where
they become semimolten and are accelerated to the workpiece at
8,000 ft/sec (2.4 km/sec). Normal deposition involves coating thick-
nesses of approximately 0.0005–0.0015 in. (0.013–0.038 mm), with
the hardness of tungsten and chromium carbide coatings at ap-
proximately 66 HRC.
Although the deposition elements are semimolten, the process
raises the work-area temperature to only about 150° F (66° C).
The coating vendor ships the coated workpieces with either a dia-
mond polish or a matte finish, depending on the customer’s re-
quirements. Preparation for coating is performed by the coating
vendor and consists of mechanical etching, that is, blasting the
surface to be coated with a special sand (MPD Company 1975).

Plasma-assisted CVD
Plasma-assisted CVD is used for the deposition of amorphous
diamond-like carbon. As in the case of PVD, high-vacuum-com-
patible equipment is used. A high-frequency (13.5 MHz) discharge

213

Ch10.pmd 213 10/29/02, 11:26 AM


Tool and Die Making Troubleshooter

is initiated. First, an argon atmosphere is used to activate the


surfaces by a sputter cleaning process. Afterward, the argon is
replaced by an atmosphere consisting of hydrocarbons and hy-
drogen. The difference between the mobility of the positive ions
and the negative electrons, and the difference of potential between
the surface of the parts to be coated, will result in a negative po-
larization of the workpieces. As a result, the ions will be acceler-
ated onto the surface to be coated.
The amount of energy applied defines the nature of the coating.
Too little energy yields soft, polymer-like coatings. If the energy is
too high, graphitic-like deposits are observed. Hence, obtaining hard,
amorphous, diamond-like carbon coatings requires close control of
the DC-potential, the partial pressures, and the gas flow. Geometri-
cal factors are also critical for optimal coating properties. There-
fore, specific fixtures must be constructed for each geometry. Since
all material to be deposited originates from the gas phase, very
even coating thicknesses are achieved. Due to the low deposition
temperature, materials such as aluminum and titanium alloys or
even plastics can be coated. The throwing power of the process is
limited only by the geometry of the discharge.
Amorphous, diamond-like carbon coatings possess high hard-
ness, a very low coefficient of friction, and corrosion resistance to
acids, alkalies, and solvents (Meyer and Bonetti Undated).

REFERENCES
Arter, Rich. 1990. “Japanese Technology Finds a Home in Indi-
ana.” Tooling and Production, October. Solon, OH: Huebcore Pub-
lishing.
Beam Alloy Corp. 1989. Ion Implantation. Dublin, OH: Beam Al-
loy Corp.
Bethlehem Steel Corp. 1980. Modern Steels & Their Properties. Hand-
book 3310. Bethlehem, PA: Bethlehem Steel Corp., pp. 66–71.
Clavel, Alfred. Undated. “Coated Tools: Two Decades of Improv-
ing Quality,” brochure. Article “Productivity CVD and PVD Coat-
ings for the Metalworking Industry.” Utica, MI: Ti-Coating, Inc.

214

Ch10.pmd 214 10/29/02, 11:26 AM


Chapter 10: Wear Enhancement Treatments

Deutchman, Arnold H. and Partyka, Robert J. 1993. “Ion-beam-


enhanced Deposition of Hard-chrome Coatings.” Aerospace Sym-
posium. Orlando, FL: American Electroplaters and Surface Fin-
ishers Society.
Dynamic Metal Treating, Inc. Undated a. Technical sales brochure,
“DYNA-BLUE Cost Effective, Increased Productivity.” Canton,
MI: Dynamic Metal Treating, Inc.
——. Undated b. Technical sales brochure, “NITROWEAR.” Can-
ton, MI: Dynamic Metal Treating, Inc.
Meyer, M. and Bonetti, R. Undated. “CVD and Plasma—CVD Pro-
tective Coatings for Machine Parts.” Beechwood, OH and Olten,
Switzerland: Sylvester & Co. and Bernex AG, respectively.
MPD (Micro Plasma Deposition) Company. 1975. “Micro-plasma
Coating—New Technology for the Metalworking Industry.” Lake
Orion, MI: MPD Company.
Paulson, Pete. 1993. “Frozen Gears.” Gear Technology, March/
April. Elk Grove Village, IL: Randal Publishing, Inc.
Reeber, R. R. and Sridharan, K. 1994. “Plasma-source Ion Im-
plantation.” Advanced Materials & Processes, December. Materi-
als Park, OH: ASM International.
Richter Precision, Inc. 1993. “State of the Art Hard Coatings.”
Modern Applications News, January. Nokomis, FL: Nelson Pub-
lishing, Inc.
Smith, James V. 1990. “Diffusion Process Extends Tool Life.”
Stamping Quarterly, October. Rockford, IL: Fabricator’s and
Manufacturer’s Association International.

215

Ch10.pmd 215 10/29/02, 11:26 AM


Ch10.pmd 216 10/29/02, 11:26 AM
Appendix A: Material Data

Appendix A:
Material Data

The data included here describe the more commonly used tool-
and-die steels mentioned throughout this book. They have been
compiled from the author’s own experiences, physical testing, and
various technical sources. While reasonable steps have been taken
to maintain their accuracy, the author does not assume responsi-
bility and/or liability for this information and suggests that read-
ers conduct their own tests to verify the data and/or seek specific
professional judgments prior to any application of these tech-
niques.

A2 TOOL STEEL
Composition: C 1.00%, Mo 1.00%, Cr 5.00%, V 0.20%.
Description: A2 is a general purpose, air-hardening tool steel.
It has a wide range of tooling applications with more abrasion
resistance than the S-series shock steels and better toughness and
ductility than the D-series wear steels.
Machinability: When properly annealed, A2 has a machin-
ability rating of 60 compared with a 1% carbon tool steel rated at
100.
Dimensional stability: When air quenched from the proper
hardening temperature, A2 can be expected to expand approxi-
mately 0.001 in./in. (0.001 mm/mm). Note that distortion (bend-
ing, bowing, or twisting) as well as part geometry can add to the
variations in movement of a part being hardened.

217

AppA.pmd 217 10/29/02, 11:27 AM


Tool and Die Making Troubleshooter

Thermal Cycling
To avoid decarburization, A2 tool steel should be annealed and/
or hardened in a controlled neutral atmosphere, vacuum, or neu-
tral-salt furnace environment.
Anneal: Heat to 1,650° F (899° C), and soak 2 hr/in. (4.7 min-
utes/mm) of thickness. Cool 40° F (22° C) per hour to 900° F (482°
C). Air cool to room temperature. Approximate annealed hard-
ness is 235 HB (maximum).
Stress relief of unhardened material: Heat slowly to 1,200–
1,250° F (649–677° C). Soak for 2 hr/in. (4.7 minutes/mm) of thick-
ness at heat. Slow cool (furnace cool if possible) to room
temperature.
Preheat: Heat to 1,200° F (649° C) and hold at this tempera-
ture until thoroughly soaked.
Harden: Heat to 1,750–1,800° F (954–982° C). Soak at heat for
45–60 minutes/in. (1.8–2.4 minutes/mm) of greatest thickness.
Quench: Air quench to 150° F (66° C). Temper immediately.
Temper: A double temper is mandatory. Soak for 2 hr/in. (4.7
minutes/mm) of thickness at heat. Slow cool to room temperature
between tempers.

Temper
°F (°C) HRC
As quenched 64
400 (204) 60
500 (260) 56
600 (316) 56
800 (427) 56
900 (482) 56
1,000 (538) 56
1,100 (593) 50
1,200 (649) 43
1,300 (704) 34
Specimens 1 in. (25.4 mm) in diameter were quenched from 1,775° F (968° C).

Stress-relief temper: A stress-relief temper for hardened


material is strongly recommended after significant grinding, weld-

218

AppA.pmd 218 10/29/02, 11:27 AM


Appendix A: Material Data

ing, or electrical discharge machining (EDM). Select a tempera-


ture that is 25 or 50° F (14 or 28° C) lower than the last tempering
temperature used (Bethlehem Steel Corp. 1981).

D2 TOOL STEEL
Composition: C 1.55%, Mo 0.80%, Cr 11.50%, V 0.90%.
Description: D2 is a high-carbon, high-chromium, air-hard-
ening tool steel formulated to combine excellent abrasion resis-
tance and air-hardening characteristics. D2 has become the
tool-and-die standard against which other tool steels are measured
for abrasion resistance, dimensional stability in hardening, and
air-hardening characteristics.
Machinability: When properly annealed, D2 has a machin-
ability rating of 45 compared to a 1% carbon steel rated at 100.
Dimensional stability: D2 has the minimum distortion in
heat treatment as compared to other tool steels. When air
quenched from the proper hardening temperature, this grade can
be expected to expand or contract approximately 0.0005 in./in.
(0.0005 mm/mm). Note that distortion (bending, bowing, or twist-
ing) as well as part geometry can add to the variations in move-
ment of a part being hardened. Refer to the information listed
under tempering for more information.

Thermal Cycling
To avoid decarburization, D2 tool steel should be annealed and/
or hardened in a controlled neutral atmosphere, vacuum, or neu-
tral-salt furnace environment.
Anneal: Heat to 1,650° F (899° C) and soak 1.5 hr/in. (3.6 min-
utes/mm) of thickness. Cool 20° F (11° C) per hour to 900° F (482°
C). Cool down in furnace to room temperature. The approximate
annealed hardness is 220 HB (maximum).
Stress relief of unhardened material: Heat slowly to 1,200–
1,250° F (649–677° C). Soak for 2 hr/in. (4.7 minutes/mm) of thick-
ness at heat. Soak and slow cool (furnace cool if possible) to room
temperature.
Preheat: Heat to 1,250° F (677° C). Hold at this temperature
until thoroughly soaked.

219

AppA.pmd 219 10/29/02, 11:27 AM


Tool and Die Making Troubleshooter

Harden: Heat to 1,850° F (1,010° C) and soak at heat for 45–60


minutes/in. (1.8–2.4 minutes/mm) of thickness. Soak the mate-
rial long enough to get all of the alloying elements into solid solu-
tion during the austenitizing cycle for proper response to heat
treatment. For items under 1 in. (25.4 mm) in thickness, soaking
time should be 45–60 minutes minimum.
Quench: Air quench to 150° F (66° C). Temper immediately.
Temper: Double tempering is mandatory, three tempers are
sometimes preferred. Soak for 2 hr/in. (4.7 minutes/mm) of thick-
ness at heat. Air cool to room temperature between tempers.
Double tempering at the range of secondary hardness, 900–960°
F (482–516° C), is strongly recommended.

Temper
°F (°C) HRC
As quenched 64
400 (204) 60
500 (260) 58
600 (316) 58
800 (427) 57
900–960 (482–516) 58–60
1,000 (538) 56
1,100 (593) 48
Specimens were air quenched from 1,850° F (1,010° C).

Stress-relief temper: A stress-relief temper for hardened


material is strongly recommended after significant grinding, weld-
ing, or EDM. Select a temperature that is 25–50° F (14–28° C)
lower than the last tempering temperature used (Bethlehem Steel
Corp. 1978a).

H13 TOOL STEEL


Composition: C 0.40%, Mo 1.00%, Cr 5.00%, V 1.00%.
Description: H13 is the most widely used steel for general hot
work applications, die casting dies, plastic-injection mold and forg-

220

AppA.pmd 220 10/29/02, 11:27 AM


Appendix A: Material Data

ing process tooling. It combines good red hardness, abrasion re-


sistance, and resistance to heat checking at hardness levels in the
range of 45–50 HRC.
Machinability: When properly annealed, H13 has a machin-
ability rating of 70 compared to a 1% carbon steel rated at 100.
Dimensional stability: When air quenched from the proper
hardening temperature, H13 tool steel can be expected to expand
approximately 0.001 in./in. (0.001 mm/mm). Note that distortion
(bending, bowing, or twisting) and part geometry can add to the
variations in movement of a part being hardened. See tempering
section.

Thermal Cycling
To avoid decarburization, H13 should be annealed and/or hard-
ened in a controlled neutral atmosphere, vacuum, or neutral-salt
furnace environment.
Anneal: Heat to 1,600° F (871° C) and soak 1 hr/in. (2.4 min-
utes/mm) of thickness. Cool 25° F (14° C) per hour to 900° F (482°
C). Air cool to room temperature. The approximate annealed hard-
ness is 220 HB (maximum).
Stress relief of unhardened material: Heat slowly to 1,200–
1,250° F (649–677° C). Soak 2 hr/in. (4.7 minutes/mm) of thickness
at heat. Slow cool (furnace cool if possible) to room temperature.
Preheat: Heat to 1,300–1,400° F (704–760° C) and hold at this
temperature until thoroughly soaked.
Harden: Heat to 1,825–1,850° F (996–1,010° C). Soak at heat
for 45–60 minutes/in. (1.8–2.4 minutes/mm) of thickness.
Quench: Air quench to 150° F (66° C). Avoid oil quenching
whenever possible. Temper immediately.
Temper: Double tempering is mandatory; three tempers are
sometimes preferred. Soak for 2 hr/in. (4.7 minutes/mm) of thick-
ness at heat. Air cool to room temperature between tempers. Die-
casting dies should be tempered to 44–48 HRC. Tooling that
requires maximum shock resistance should be tempered to 40–44
HRC.

221

AppA.pmd 221 10/29/02, 11:27 AM


Tool and Die Making Troubleshooter

Temper
°F (°C) HRC
As quenched 52–54
400 (204) 52–54
500 (260) 53
600 (316) 53
800 (427) 53
900 (482) 52–54
1,000 (538) 52–54
1,050 (566) 48–50
1,100 (593) 46–48
1,150 (621) 40–44
Specimens 1 in. (25.4 mm) in diameter were air quenched from 1,850° F (1,010° C).

Stress-relief temper: A stress-relief temper for hardened


material is strongly recommended after significant grinding, weld-
ing, or EDM. Select a temperature that is 25 or 50° F (14 or 28° C)
lower than the last tempering temperature used (Bethlehem Steel
Corp. 1978b).

M2 HIGH-SPEED STEEL
Composition: C 0.83%, Cr 4.15%, Mo 5.00%, V 1.90%, W 6.35%.
Description: M2 is a general-purpose, moly high-speed steel
characterized by a balanced combination of abrasion resistance,
toughness, and good red hardness.
Machinability: When properly annealed, M2 has a machin-
ability rating of 65 compared to a 1% carbon steel rated at 100.
Dimensional stability: When air quenched from the proper
hardening temperature, M2 can be expected to expand approxi-
mately 0.001 in./in. (0.001 mm/mm). Note that distortion (bend-
ing, bowing, or twisting) as well as part geometry can add to the
variations in movement of a part that is being hardened.

Thermal Cycling
To avoid decarburization, M2 should be annealed and/or hard-
ened in a controlled neutral atmosphere, vacuum, or neutral-salt
furnace environment.

222

AppA.pmd 222 10/29/02, 11:27 AM


Appendix A: Material Data

Anneal: Heat to 1,600° F (871° C) and soak thoroughly at heat.


Furnace cool 25° F (14° C) per hour to 900° F (482° C). Air cool to
room temperature. Approximate annealed hardness is 241 HB
(maximum).
Stress relief of unhardened material: Heat slowly to 1,200–
1,250° F (649–677° C). Soak 2 hr/in. (4.7 minutes/mm) of thickness
at heat. Slow cool (furnace cool if possible) to room temperature.
Preheat: Heat slowly to 1,550° F (843° C), hold until thoroughly
soaked, heat to 1,850° F (1,010° C), and soak thoroughly.
Harden: Total heating time in the furnace varies from a few
minutes to a maximum of 15 minutes, depending on the size of
the tool, the relative heat capacity of the furnace, and the size of
the charge. Heat to 2,150–2,200° F (1,177–1,204° C) for maximum
toughness and minimum distortion. Heat to 2,250–2,275° F (1,232–
1,246° C) for maximum hardness and abrasion resistance.
Quench: To develop full hardness, oil quenching to 150–200°
F (66–93° C) is preferred. For air quenching, quench to 150° F (66°
C). For salt quenching, the hot salt should be maintained just above
the Ms temperature. After equalizing, parts should be withdrawn
from the salt and air cooled to 150° F (66° C). Temper immediately.
Temper: Double tempering is mandatory; three tempers are
sometimes preferred. Soak for 2 hr/in. (4.7 minutes/mm) of thick-
ness. Air cool to room temperature between tempers. The best
tempering range for M2 is 1,000–1,050° F (538–566° C). This re-
sults in the best combination of cutting ability, hardness, strength,
and toughness.
Temper
°F (°C) HRC
As quenched 64
400 (204) 63
500 (260) 62.5
600 (316) 62.5
700 (371) 62.5
800 (427) 63.5
900 (482) 64
1,000 (538) 65.5
1,050 (566) 63.5
1,100 (593) 61.5
1,150 (621) 60
1,200 (649) 53
Specimens 1 in. (25.4 mm) in diameter were air quenched from 2,250° F (1,232° C).

223

AppA.pmd 223 10/29/02, 11:27 AM


Tool and Die Making Troubleshooter

Stress-relief temper: A stress-relief temper for hardened


material is strongly recommended after significant grinding, weld-
ing, or EDM. Select a temperature that is 25 or 50° F (14 or 28° C)
lower than the last tempering temperature used (Bethlehem Steel
Corp. 1978c).

M4 HIGH-SPEED STEEL
Composition: C 1.27%, Cr 4.50%, Mo 4.50%, V 4.00%, W 5.55%.
Description: M4 is a moly-tungsten, high-speed steel with high
carbon and vanadium content. Its resistance to abrasion is supe-
rior to most high-speed steels, and it is the material of choice for
broaches, reamers, checking tools, blanking dies, swaging dies,
and punches for abrasive materials.
Machinability: When properly annealed, M4 has a machin-
ability rating of 40 compared to a 1% carbon steel rated at 100.
Dimensional stability: When air quenched from the proper
hardening temperature, M4 can be expected to expand approxi-
mately 0.001 in./in. (0.001 mm/mm). Note that distortion (bend-
ing, bowing, or twisting) as well as part geometry can add to the
variations in movement of a part that is being hardened.

Thermal Cycling
To avoid decarburization, M4 should be annealed and/or hard-
ened in a controlled neutral atmosphere, vacuum, or neutral-salt
furnace environment.
Anneal: Heat to 1,600° F (871° C) and soak thoroughly at heat.
Furnace cool 25° F (14° C) per hour to 900° F (482° C) and air cool
to room temperature. The approximate annealed hardness is 241
HB (maximum).
Stress relief of unhardened material: Heat slowly to
1,200–1,250° F (649–677° C). Soak for 2 hr/in. (4.7 minutes/mm)
of thickness at heat. Slow cool (furnace cool if possible) to room
temperature.
Preheat: Heat to 1,550° F (843° C), hold until thoroughly soaked,
heat to 1,850° F (1,010° C), and hold until thoroughly soaked.

224

AppA.pmd 224 10/29/02, 11:27 AM


Appendix A: Material Data

Harden: Heat rapidly to 2,150–2,250° F (1,177–1,232° C). To-


tal heating time in the furnace varies from a few minutes to a
maximum of 15 minutes, depending of the size of the tool. Heat to
2,150° F (1,177° C) for maximum toughness and minimum distor-
tion. Heat to 2,250° F (1,232° C) for maximum hardness and abra-
sion resistance.
Quench: M4 may be oil quenched to 150–200° F (66–93° C), or
air hardened to 150° F or quenched in salt. For salt quenching,
the hot salt should be maintained just above the Ms temperature.
After equalizing, the part should be withdrawn from the salt and
air cooled to 150° F (66° C). Temper immediately after the quench.
Temper: Double tempering is mandatory; three tempers are
preferred. Soak for 2 hr/in. (4.7 minutes/mm) of thickness at heat.
Air cool to room temperature between tempers. The optimum
range of hardness for most tooling applications involving M4 is
58–62 HRC.

Temper
°F (°C) HRC
As quenched 64
400 (204) 62.5
500 (260) 61.0
600 (316) 60.5
700 (371) 61.5
800 (427) 62.0
900 (482) 64
1,000 (538) 66
1,100 (593) 65
1,150 (621) 63
1,200 (649) 55
Specimens 0.75 in.2 × 2 in. (48.4 cm2 × 5.1 cm) long were quenched in air from
2,250° F (1,232° C).

Stress-relief temper: A stress-relief temper for hardened


materials is strongly recommended after significant grinding,
welding, or EDM. Select a temperature that is 25–50° F (14–28°
C) lower than the last tempering temperature used (Teledyne Vasco
1968).

225

AppA.pmd 225 10/29/02, 11:27 AM


Tool and Die Making Troubleshooter

O1 TOOL STEEL
Composition: C 0.90%, Mn 1.20%, V 0.20%, W 0.50%, Cr 0.50%.
Description: O1 tool steel is a general-purpose, oil-harden-
ing, tool-and-die steel. Normal care in heat treatment gives good
results in hardening and produces small dimensional changes. O1
has good abrasion resistance and sufficient toughness for a wide
variety of tooling applications.
Machinability: When properly annealed, O1 has a machin-
ability rating of 90 compared to a 1% carbon steel rated at 100.
Dimensional stability: When oil quenched from the proper
hardening temperature, O1 can be expected to expand approxi-
mately 0.0015 in./in. (0.0015 mm/mm). Note that distortion (bend-
ing, bowing, or twisting) and part geometry can add to the
variations in movement of a part being hardened.

Thermal Cycling
To avoid decarburization, O1 should be annealed and/or hard-
ened in a controlled neutral atmosphere, vacuum, or neutral-salt
furnace.
Anneal: Heat slowly to 1,450° F (788° C) and soak thoroughly.
Cool 25° F (14° C) per hour to 900° F (482° C). Air cool to room
temperature. The approximate annealed hardness is 221 HB
(maximum).
Stress relief of unhardened material: Heat slowly to 1,250°
F (677° C). Soak for 2 hr/in. (4.7 minutes/mm) of thickness at
heat. Slow cool (furnace cool if possible) to room temperature.
Preheat: Heat to 1,200° F (649° C) and hold at this tempera-
ture until thoroughly soaked.
Harden: Heat to 1,475–1,500° F (802–816° C). Soak at heat for
30 minutes/in. (1.2 minutes/mm) of thickness.
Quench: Oil quench to 150–200° F (66–93° C). Temper imme-
diately.
Temper: Normally, oil-hardening steels need to be single tem-
pered only. However, double tempering may sometimes be pre-
ferred. Soak at heat for 2 hr/in. (4.7 minutes/mm) of thickness for
each temper. Air cool to room temperature between tempers. The
normal tempering range for O1 is 300–450° F (149–232° C).

226

AppA.pmd 226 10/29/02, 11:27 AM


Appendix A: Material Data

Temper
°F (°C) HRC
As quenched 64–65
350 (177) 62–63
400 (204) 62
500 (260) 60
600 (316) 57
700 (371) 53
800 (427) 50
900 (482) 47
1,000 (538) 44
1,100 (593) 39
Specimens 1 in. (25.4 mm) in diameter were oil quenched from 1,475° F (802° C).

Stress-relief temper: A stress-relief temper for hardened


material is strongly recommended after significant grinding, weld-
ing, or EDM. Select a temperature that is 25 or 50° F (14 or 28° C)
lower than the last tempering temperature used (Bethlehem Steel
Corp. 1977).

S5 TOOL STEEL
Composition: C 0.60%, Si 1.85%, Mn 0.70%, Mo 0.45%, V
0.20%.
Description: S5 is a silicon-manganese, moly-vanadium, shock-
resisting tool steel. It was developed primarily for shock-resisting
parts in which a combination of good ductility and hardness is
required. S5 is primarily an oil-hardening steel. Intricate parts
made from this grade should always be oil quenched.
Machinability: When properly annealed, S5 has a machin-
ability rating of 60 compared to 1% carbon steel rated at 100.
Dimensional stability: When oil quenched from the proper
hardening temperature, S5 can be expected to expand approxi-
mately 0.0015 in./in. (0.0015 mm/mm). Note that distortion (bend-
ing, bowing, or twisting) and part geometry can add to the
variations in movement of a part being hardened.

227

AppA.pmd 227 10/29/02, 11:27 AM


Tool and Die Making Troubleshooter

Thermal Cycling
To avoid decarburization, S5 should be annealed and/or hard-
ened in a controlled neutral atmosphere, vacuum, or neutral-salt
furnace.
Anneal: Heat slowly to 1,450° F (788° C) and soak 1 hr/in. (2.4
minutes/mm) of thickness. Cool 25° F (14° C) per hour to 900° F
(482° C). Air cool to room temperature. The approximate annealed
hardness is 230 HB (maximum).
Stress relief of unhardened material: Heat slowly to 1,250°
F (677° C). Soak for 2 hr/in. (4.7 minutes/mm) of thickness at
heat. Slow cool (furnace cool if possible) to room temperature.
Preheat: Heat to 1,250° F (677° C) and hold at this tempera-
ture until thoroughly soaked.
Harden: Heat to 1,600° F (871° C) and soak at heat for 30 min-
utes/in. (1.2 minutes/mm) of thickness.
Quench: Oil quench to 150° F (66° C). Temper immediately.
Temper: Normally, oil-hardening steels need be single tempered
only. However, double tempering may sometimes be preferred. Soak
at heat for 2 hr/in. (4.7 minutes/mm) of thickness for each tem-
per. Air cool to room temperature between tempers. The normal
tempering range for S5 is 400–650° F (204–343° C).

Temper
°F (°C) HRC
As quenched 63–64
400 (204) 60–62
500 (260) 60–61
600 (316) 59–60
700 (371) 57–58
800 (427) 52–53
900 (482) 50–51
1,000 (538) 48–49
1,100 (593) 46–47
1,200 (649) 40–41
Specimens 0.75 in. (19.1 mm) in diameter were oil quenched from 1,600° F
(871° C).

Stress-relief temper: A stress-relief temper for hardened


material is strongly recommended after significant grinding, weld-

228

AppA.pmd 228 10/29/02, 11:27 AM


Appendix A: Material Data

ing, or EDM. Select a temperature that is 25 or 50° F (14 or 28° C)


lower than the last tempering temperature used (Bethlehem Steel
Corp. 1978d).

S7 TOOL STEEL
Composition: C 0.50%, Si 0.25%, Mn 0.70%, Cr 3.25%, Mo
1.40%.
Description: S7 is a shock-resisting tool steel with excellent
impact properties. Because S7 is an air-hardening steel, it is safe
and stable in heat treatment. Its most important characteristic is
its versatility. S7 is used widely for medium cold-work tools and
dies, plastic-molding dies, shear blades, medium hot-work dies,
master hobs, and component parts of many products.
Machinability: When properly annealed, S7 has a machin-
ability rating of 70 compared to a 1% carbon steel rated at 100.
Dimensional stability: When air quenched from the proper
hardening temperature, S7 can be expected to expand approxi-
mately 0.001 in./in. (0.001 mm/mm). Note that distortion (bend-
ing, bowing, or twisting) and part geometry can add to the
variations in movement of a part being hardened.

Thermal Cycling
To avoid decarburization, S7 should be annealed and/or hard-
ened in a controlled atmosphere, vacuum, or neutral-salt furnace
environment.
Anneal: Heat to 1,550° F (843° C) and soak for 1.5 hr/in. (3.5
minutes/mm) of thickness. Cool 25° F (14° C) per hour to 900° F
(482° C). Air cool to room temperature. The approximate annealed
hardness is 230 HB (maximum).
Stress relief of unhardened material: Heat slowly to 1,250°
F (677° C). Soak for 2 hr/in. (4.7 minutes/mm) of thickness at
heat. Slow cool (furnace cool if possible) to room temperature.
Preheat: Heat to 1,250° F (677° C) and hold at this tempera-
ture until thoroughly soaked.
Harden: Heat to 1,725–1,750° F (941–954° C) and soak at heat
for 45–60 minutes/in. (1.8–2.4 minutes/mm) of thickness. Sizes
under 1 in. (25.4 mm) thick should be held for 45–60 minutes
minimum.

229

AppA.pmd 229 10/29/02, 11:27 AM


Tool and Die Making Troubleshooter

Quench: Air quench sections up to 2.50 in. (63.5 mm) thick in


still air. Quench to 150° F (66° C). Temper immediately after
quenching in all cases.
Temper: Double tempering is mandatory. Soak for 2 hr/in. (4.7
minutes/mm) of thickness for each temper. Air cool to room tem-
perature between tempers. For cold-work applications, the nor-
mal tempering range is 400–500° F (204–260° C). For hot-work
applications, a tempering temperature of 900–1,000° F (482–538°
C) is suggested. Never temper S7 under 400° F (204° C).

Temper
°F (°C) HRC
As quenched 58–60
400 (204) 58
500 (260) 56
600 (316) 55
700 (371) 54
800 (427) 53
900 (482) 52
1,000 (538) 51
1,100 (593) 47
1,200 (649) 38
Specimens 1 in. (25.4 mm) in diameter and 3 in. (76.2 mm) long were air
hardened from 1,725° F (941° C).

Stress-relief temper: A stress-relief temper for hardened


material is strongly recommended after significant grinding, weld-
ing, or EDM. Select a temperature that is 25 or 50° F (14 or 28° C)
lower than the last tempering temperature used (Bethlehem Steel
Corp. 1976).

W1 CARBON TOOL STEEL


Composition: Carbon 0.70/1.30%.
Description: W1 is a cold-work, water-hardening tool steel. This
steel grade depends on its relatively high carbon content for its
useful properties. W1 is a very versatile grade noted for its easy
machining characteristic, ability to develop a keen cutting edge,
high surface-case hardness, and soft, ductile, tough inner core.

230

AppA.pmd 230 10/29/02, 11:27 AM


Appendix A: Material Data

Machinability: When properly annealed, W1 has a machin-


ability rating of 100 compared to a 1% carbon steel rated at 100.
Dimensional stability: When water quenched from the proper
hardening temperature, W1 can be expected to expand approxi-
mately 0.002–0.004 in./in. (0.002–0.004 mm/mm). Note that dis-
tortion (bending, bowing, or twisting) as well as part geometry
can add to the variations in movement of a part being hardened.

Thermal Cycling
To avoid decarburization, W1 should be annealed and/or hard-
ened in a controlled neutral atmosphere, vacuum, or neutral-salt
furnace environment.
Anneal: Heat to 1,375–1,400° F (746–760° C) and soak until
uniformly heated, approximately 30 minutes/in. (1.2 minutes/mm).
Furnace cool 50° F (28° C) per hour to 975° F (524° C) and air cool
to room temperature. The approximate annealed hardness is 200
HB (maximum).
Stress relief of unhardened material: Heat slowly to 1,200–
1,250° F (649–677° C). Soak 2 hr/in. (4.7 minutes/mm) of thick-
ness at heat. Soak and slow cool (furnace cool if possible) to room
temperature.
Preheat: Heat to 1,200° F (649° C) and hold at this tempera-
ture until thoroughly soaked.
Harden: Heat to 1,425–1,475° F (774–802° C). Soak at heat for
30 minutes/in. (1.2 minutes/mm) of thickness. Temperatures on
the higher side of this range will increase the depth of the case or
chill. Use the low end of the range for small sizes, and the higher
end of the range for larger sizes.
Quench: W1 may be water quenched, but brine quenching
is preferred. Water or brine quench to 150–200° F (66–93° C).
Oil quenching is sometimes used for light sections and where
maximum hardness is not required. Temper immediately after
quenching.
Temper: Normally, water-hardening steels need to be single
tempered only. However, double tempering may sometimes be pre-
ferred. Soak at heat for 2 hr/in. (4.7 minutes/mm) of thickness for
each temper. Air cool to room temperature between tempers.

231

AppA.pmd 231 10/29/02, 11:27 AM


Tool and Die Making Troubleshooter

Water Quench FFrom


rom Depth of Chill Tempering
°F (°C) 64ths/in. (mm) °F (°C) HRC
1,375 (746) 6.0 (2.38) As-quenched 67
1,400 (760) 7.0 (2.78) 300 (149) 64
1,425 (774) 8.0 (3.18) 400 (204) 61
1,450 (788) 8.5 (3.37) 500 (260) 59
1,450 (788) 8.5 (3.37) 600 (316) 55
Specimens 0.75 in. (19.1 mm) in diameter and 3 in. (76.2 mm) long were water
quenched from 1,450° F (788° C).

Stress-relief temper: A stress-relief temper for hardened


material is strongly recommended after significant grinding, weld-
ing, or EDM. Select a temperature that is 25 or 50° F (14 or 28° C)
lower than the last tempering temperature used (Bethlehem Steel
Corp. Undated).

420 MARTENSITIC STAINLESS STEEL


Composition: C 0.38%, Mn 0.40%, Si 0.40%, Cr 13.00%.
Description: 420 stainless steel is a hardenable stainless steel.
This stainless alloy has an excellent combination of corrosion re-
sistance, abrasion resistance, and polishability with a working
hardness of 46–52 HRC.
Machinability: When properly annealed, 420 has a machin-
ability rating of 70 compared to a 1% carbon steel rated at 100.

Thermal Cycling
To avoid decarburization, 420 should be annealed and/or hard-
ened in a controlled neutral atmosphere, vacuum, or neutral-salt
furnace environment.
Anneal: Heat to 1,650° F (899° C), soak, and uniformly heat
throughout. Cool slowly with furnace to 600° F (316° C), then air
cool to room temperature. The approximate annealed hardness is
230 HB (maximum).
Stress relief of unhardened material: Heat slowly to
1,200–1,250° F (649–677° C). Soak 2 hr/in. (4.7 minutes/mm) of
thickness at heat. Slow cool (furnace cool if possible) to room tem-
perature.

232

AppA.pmd 232 10/29/02, 11:27 AM


Appendix A: Material Data

Preheat: Heat to 1,250–1,300° F (677–704° C), holding at this


temperature until thoroughly soaked.
Harden: Heat to 1,750–1,850° F (954–1,010° C). Soak at heat
for 30 minutes/in. (1.2 minutes/mm) of thickness. If maximum
hardness and corrosion resistance is needed, use the high side of
the hardening temperature.
Quench: 420 can be oil or air quenched. In complex or irregu-
lar sections, use air quenching for less distortion and greater as-
surance of freedom from cracking. If higher hardness is desired
and the part configuration permits, use oil quenching. Quench in
air to 150° F (66° C). Oil quench to 150–200° F (66–93° C). In
either case, temper immediately after quenching.
Temper: Double tempering is mandatory; three tempers are
sometimes preferred. Soak for 2 hr/in. (4.7 minutes/mm) of thick-
ness at heat for each tempering cycle. Air cool to room tempera-
ture between tempers. For most applications, temper 420 stainless
at 400° F (204° C) minimum. Avoid tempering above 800° F (427°
C) because a drop in impact strength and corrosion resistance
will result, a condition that disappears when the tempering tem-
perature is 1,100° F (593° C) or higher. Tempering temperatures
of 1,100° F (593° C) or higher increase toughness at a sacrifice in
hardness.

Temper HRC Harden to


°F (°C) 1,850° F (1,010° C)
As quenched 52–54
400 (204) 51–53
500 (260) 51–53
600 (316) 51–53
1,000 (538) 47–48
1,050 (566) 44–46
1,100 (593) 41–43
1,150 (621) 36–38
Specimens 1 in. (25.4 mm) in diameter were quenched in oil from 1,800° F (982° C).

Stress-relief temper: A stress-relief temper for hardened


material is strongly recommended after significant grinding, weld-
ing, or EDM. Select a temperature that is 25–50° F (14–28° C)

233

AppA.pmd 233 10/29/02, 11:27 AM


Tool and Die Making Troubleshooter

lower than the last tempering temperature used (Bethlehem Steel


Corp. 1978e).

REFERENCES
Bethlehem Steel Corp. October 1976. “Bearcat AISI S7 Bethlehem
Shock-resisting Tool Steel.” Folder 2412-B. Bethlehem, PA:
Bethlehem Steel Corp.
——. 1977. “BTR AISI Type O1 Bethlehem Cold-work Tool Steel.”
Folder 2321-B. Bethlehem, PA: Bethlehem Steel Corp.
——. 1978a. “Lehigh H AISI Type D2 Bethlehem Cold-work Tool
Steel.” Folder 2323-B. Bethlehem, PA: Bethlehem Steel Corp.
——. 1978b. “Cromo-high V AISI Type H13 Bethlehem Hot-work
Tool Steel.” Folder 2133-C. Bethlehem, PA: Bethlehem Steel Corp.
——. 1978c. “AISI Type M2 High-speed Steel.” Folder 2484-B.
Bethlehem, PA: Bethlehem Steel Corp.
——. 1978d. “Omega AISI Type S5 Bethlehem Shock-resisting
Tool Steel.” Folder 2151-E. Bethlehem, PA: Bethlehem Steel Corp.
——. 1978e. “Bethadur 420 AISI Type 420 Bethlehem Plastic-
molding Tool Steel.” Folder 3040-A. Bethlehem, PA: Bethlehem
Steel Corp.
——. 1981. “A-H5 AISI Type A2 Bethlehem Cold-work Tool Steel.”
Folder 2322-D. Bethlehem, PA: Bethlehem Steel Corp.
——. Undated. “Carbon AISI Type W1 Bethlehem Water-harden-
ing Tool Steel.” Booklet 2673. Bethlehem, PA: Bethlehem Steel
Corp.
Teledyne Vasco. 1968. “High-speed Steels.” Latrobe, PA: Teledyne
Vasco.

234

AppA.pmd 234 10/29/02, 11:27 AM


Appendix B: Glossary

Appendix B: Glossary

A
abrasion: The loss of surface material through rubbing, grind-
ing, or frictional wear.
abrasive wear: The cutting and scratching of the surface of a
metal when abrasive particles rub against it or between two
surfaces.
Ac1: For irons and steels, the temperature at which austenite be-
gins to form on heating.
Ac3: The temperature at which transformation of ferrite to aus-
tenite is completed on heating.
adhesive wear: Surface damage most commonly in the form of
grooving caused by metal-to-metal contact (usually under heavy
compressive loads, high pressures, and/or with the development
of frictional heat). This type of metal-to-metal contact results
in welding and the breaking of these welds. When this happens
in a highly concentrated area, it is called galling and/or metal
pickup.
age hardening: A process of aging that increases hardness and
strength (tensile properties), and ordinarily decreases ductil-
ity. Age hardening usually follows rapid cooling or cold work-
ing.
aging: The terms aging, age hardening, and precipitation hard-
ening are synonymous. They refer to a type of hardening and

235

AppB.pmd 235 10/29/02, 11:27 AM


Tool and Die Making Troubleshooter

strengthening of an alloy that may occur after hot working,


heat treatment (natural or artificial aging), or cold working
(strain aging). Aging usually involves reheating a solution-an-
nealed material to a temperature lower than the solution treat-
ing temperature, followed by rapid cooling to room temperature.
This change in properties is often, but not always, due to a phase
change (precipitation of constituents for a supersaturated solid
solution). It never involves a change in chemical composition of
the metal or alloy.
air cooling: Cooling a heated metal or workpiece in still or forced
air (fan cooled).
aircraft quality: Steels intended for important or highly stressed
parts and components in the aerospace industry. They are also
used in non-aerospace applications where this quality level is
deemed necessary. Special steelmaking practices, more rigid
inspection techniques, and more restrictive selection are nec-
essary to meet the rigid aircraft quality standards.
air-hardening steel: A steel with enough carbon and other al-
loying elements to fully harden with an air quench.
alloy: A substance made up of two or more metals, or of a metal
and a nonmetal.
alloying element: An element added to a metal to effect changes
in the mechanical or physical properties of the metal. Chro-
mium, manganese, vanadium, molybdenum, tungsten, and ti-
tanium are examples of alloying elements added to iron in
steelmaking.
alloy steel: Steel containing substantial quantities of alloying
elements—other than carbon and the commonly accepted
amounts of manganese, sulfur, silicon, and phosphorus—for the
purpose of increased hardness, strength, or chemical resistance.
Alloy steels are usually considered as those containing a total
of less than 9% of such added constituents.
ambient temperature: Normal temperature of the surround-
ing area, such as shop temperature.
annealing: A term describing a number of different treatments
to heat steel and hold it at a suitable temperature followed by
slow cooling. It is used primarily to soften metallic materials

236

AppB.pmd 236 10/29/02, 11:27 AM


Appendix B: Glossary

and improve microstructure and other properties, such as ma-


chinability and cold-working characteristics.
arc welding: Welding accomplished by using an electric arc
formed between a metal or carbon electrode and the metal be-
ing welded; between two separate electrodes, as in atomic hy-
drogen welding; or between two separate pieces being welded,
as in flash welding.
artificial aging: Aging at a temperature higher than ambient or
room temperature.
ASTM: American Society for Testing and Materials, an organiza-
tion that issues standard specifications for materials, including
metals and alloys.
austempering: A thermal treatment process that involves
quenching steel from a temperature above the transformation
range into a medium having a rate of heat extraction high
enough to prevent the formation of high-temperature transfor-
mation products. The material is then held at this tempera-
ture, which is above that of martensite formation, until
transformation is complete. The product formed is termed
“lower bainite.”
austenite: A metallurgical term referring to a solid solution of
one or more elements in face-centered-cubic iron. Tool steels
must be transformed to austenite before they can be quenched
to hard martensite. Austenite is nonmagnetic.
austenitizing: The process of heating a ferrous alloy into its
transformation range to partially form austenite, or above the
transformation range to completely transform the microstruc-
ture to austenite.

B
bainite: A transformation product of austenite that forms at tem-
peratures lower than where fine pearlite is formed and tem-
peratures higher than where martensite is formed.
base metal: (1) The metal in highest proportion in an alloy. (2)
Workpiece metal that is not affected by a process like heat

237

AppB.pmd 237 10/29/02, 11:27 AM


Tool and Die Making Troubleshooter

treating, flame hardening, or welding a substrate or surround-


ing area.
BHN: Abbreviation for Brinell hardness number.
billet: An intermediate round or square produced from hot work-
ing an ingot. Billets are further reduced by forging or rolling
into bars.
bloom: Semifinished products, hot-rolled from ingots or cast from
a continuous caster machine.
blowhole: A hole in a casting or weld caused by gas entrapped
during solidification.
brazing: Joining metals by fusion of nonferrous alloys that have
melting points above 800° F (427° C), but lower than those of
the metals being joined. This may be accomplished by means of
a torch (torch brazing), in a furnace (furnace brazing), or by
dipping in a molten flux bath (dip or flux brazing).
Brinell hardness test: A method used for measuring the hard-
ness of metals. A smooth, ground, flat spot on the surface of the
metal is subjected to indentation by a hardened steel ball un-
der a specific pressure load. A special microscope with a cali-
brated lens is then used to measure the indentation diameter.
The diameter is referenced on a special chart to allow for direct
determination of the Brinell hardness.
brittleness: A tendency to fracture without appreciable defor-
mation.
BTU: British thermal unit—the amount of heat needed to increase
the temperature of 1 lb (0.45 kg) of water 1° F (–17° C).
burning: Heating a metal to the point of incipient melting or
where intergranular oxidation begins.
burnishing: To polish and/or make smooth by rubbing. Suffi-
cient friction may be caused by the rubbing action to discolor/
tint (turn straw brown) the surface of the workpiece. Stresses
caused by burnishing during the polishing action may some-
times lead to orange peel.
burnt: A term applied to a metal permanently damaged by over-
heating.
burr: A thin ridge or roughness left by a cutting, shearing, or
sawing operation.
butt welding: Joining two edges by placing one against the other
and welding them.

238

AppB.pmd 238 10/29/02, 11:27 AM


Appendix B: Glossary

C
camber: A bend in a plate or other product of a rolling mill that
results when one edge or side is longer than the other. Camber
in plates is often caused by rolls that are closer together at one
end than at the other, or by uneven temperatures in the slab.
In rails and structural shapes, the camber is the up or down
curvature, as distinguished from the sidewise curvature or
sweep.
carbide: A particle formed by the combination of carbon with
one or more metallic elements. Chromium, molybdenum, tung-
sten, titanium, and vanadium are common carbide-forming el-
ements.
carbon steel: Steel that owes its properties chiefly to various
percentages of carbon without substantial amounts of other
alloying elements; also known as straight carbon steel, or plain
carbon steel. Steel is classified as carbon steel when no mini-
mum content or element other than carbon is specified or re-
quired to obtain a desired alloying effect; when the specified
minimum for copper does not exceed 0.40%; or when the fol-
lowing alloying elements are not exceeded: manganese 1.65%,
silicon 0.60%, or copper 0.60%.
carbonitriding: A process that implants carbon and nitrogen
into a steel’s surface by heating in a gaseous atmosphere con-
taining hydrocarbons and ammonia.
carburizing: A process that brings an austenitized ferrous ma-
terial into contact with a carbonaceous atmosphere or medium
of sufficient carbon potential to cause absorption of carbon at
the surface and create by diffusion a concentration gradient.
Hardening by quenching follows.
case: A surface layer of a ferrous alloy made harder than its inte-
rior or core.
case hardening: One or more processes for hardening steel in
which the outer portion or case is made substantially harder
than the inner portion or core. Most of the processes involve
enriching the surface layer with carbon and/or nitrogen, usu-
ally followed by quenching and tempering, or the selective hard-
ening of the surface layer by means of flame or induction
hardening.

239

AppB.pmd 239 10/29/02, 11:28 AM


Tool and Die Making Troubleshooter

cast iron: Like steel, cast iron identifies a large family of ferrous
alloys, usually alloys of iron that contain 2% carbon and 1–3%
silicon. There are many different grades of cast irons contain-
ing magnesium, chromium, molybdenum, nickel, and copper.
cementite: A hard, brittle compound of iron and carbon (Fe3C)—
the major form in which carbon occurs in steel.
checks, heat checks: Numerous, very fine cracks at the surface
of a metal part. Causes may include abusive grinding, grinding
of an improperly tempered part, thermal fatigue from rapid
cyclic heating and cooling, etc.
chrome plating: An electrolytic deposit of a hard chromium layer
on the surface of tools and dies to improve abrasive wear, lu-
bricity, etc.
coated electrode: An arc-welding metal electrode lightly coated
with metal oxides and silicates for stabilizing the arc.
coating: The process of covering steel with another material,
primarily for wear and/or corrosion resistance.
coefficient of thermal expansion: The expected degree of ex-
pansion of a material per degree of temperature change, ex-
pressed as µin./in./°F (µm/mm/°C). Coefficient of expansion data
are applicable to specific temperature ranges for which the co-
efficient was developed.
coining: A stamping process used to sharpen or change an exist-
ing radius or profile, or to produce a well-defined imprint from
the die onto the workpiece.
combined carbon: The percentage of carbon in steel or cast iron
present other than as free carbon (graphite).
compressive strength: The maximum compressive stress that
a material is capable of withstanding under a squeezing or com-
pressive load.
cooling stresses: Residual stresses resulting from differential
cooling rates within a part being cooled.
core: The interior portion of a steel where a soft, ductile struc-
ture exists, as opposed to the outside case, which is harder.
corrosion: Gradual chemical or electromechanical attack on a
metal by atmosphere, moisture, or other agents.
coupling distance: The recommended distance between two
objects or forces (for example, the optimum distance between a
flame and the workpiece during the flame-hardening operation).

240

AppB.pmd 240 10/29/02, 11:28 AM


Appendix B: Glossary

crater cracks: Cracks across the center of a weld bead resulting


from heat shrinkage.
critical cooling rate: The minimum rate at which a steel can be
quenched to develop martensite and fully harden.
critical points: Temperatures at which phase changes take place
within a metal either upon heating or cooling.
crocus cloth: A very fine, abrasive polishing cloth.
cross grinding: A grinding technique used to improve surface
finish by alternating successive grinding passes at 90° angles,
each with a finer grit.
crown: In plates, sheets, or strips, the crown is characterized by
a greater thickness in the middle than at the edges. It may be
caused by deflection (bending) of the rolls or by worn rolls. The
latter is sometimes called “hollowness of the mill.”
cryogenic quench: The freezing of hardened tooling at a tem-
perature of approximately –320° F (–196° C).
crystal: A solid with its atoms arranged in a regular three-di-
mensional pattern.

D
deburring: A method for removing sharp or raw metal edges by
rolling parts in a barrel with abrasives, or by filing, grinding, or
hammering them flat and smooth.
decarburization: Carbon depletion or loss from the surface of a
carbon-containing alloy. See oxidation.
deep freeze: The freezing of hardened tooling at approximately
–120° F (–84° C).
deformation: The amount a material permanently increases or
decreases in length under load.
dendrite: A tree-like branching crystal structure that forms when
a cast metal is allowed to slowly solidify.
deoxidizing: The removal of oxygen from molten metal.
deposition rate: The speed at which filler metal is added to a
weld joint, usually in terms of volume of metal deposited per
minute.
die sinking: Using a machining process to create a depressed
pattern or cavity in a forging die.

241

AppB.pmd 241 10/29/02, 11:28 AM


Tool and Die Making Troubleshooter

differential hardening: Any hardening procedure that produces


a significant variation of properties in the same piece.
dirt: See inclusions.
double tempering: Performing two complete tempering cycles
after austenitizing and quenching heat treatments.
drawing: Another term used to describe tempering.
drawing quality: Deep-drawing rimmed steel or extra-deep-
drawing aluminum-killed steels for rolling and processing op-
erations that can stand extreme pressing, drawing, or forming
forces without creating defects.
drop forging: Closed-die forging using dies designed to produce
the required shape, usually by impact rather than pressing.
ductile cast iron: One of the four categories of cast iron, ductile
cast iron has superior properties to gray and white cast iron. It
is also called “nodular cast iron.”
ductility: A measure of a material’s ability to stretch plastically
without fracturing.
dwell: A term used during the flame-hardening operation to de-
scribe the travel speed of the torch flame as it heats the
workpiece.

E
elastic limit: The maximum stress to which a material can be
subjected without permanent deformation.
elasticity: The ability of a material to return to its original shape
and dimensions after a deforming load has been removed.
electrical conductivity: A measure of a material’s ability to
conduct electrical current.
electrical resistance: The resistance of a material to the flow
of electrical current.
embrittlement: The loss of a metal’s ductility due to changes in
chemistry, increased hardness, and/or microstructural changes.
emery cloth: An abrasive cloth, available in a variety of grits,
used for grinding and polishing.
extrusion: A metal-forming process that forces a metal slug
through a die to produce a particular shape.

242

AppB.pmd 242 10/29/02, 11:28 AM


Appendix B: Glossary

F
face-centered cubic: The crystal structure of austenite.
ferrite: Pure iron of body-centered-crystal structure. It is soft
and ductile.
filler metal: Metal added in the making of a brazed, soldered, or
welded joint.
fillet: The curvature at the intersection of two perpendicular sur-
faces.
finishing stone: An abrasive stone compound used to smooth
and/or polish the surface of a part.
firebox quality: Plates for pressure vessels capable of being ex-
posed to fire or heat as well as thermal and mechanical stresses.
flame hardening: Hardening by directing a hot flame onto the
surface of a ferrous alloy to heat it above the critical tempera-
ture (that is, into the austenitic temperature range) and then
rapidly cooling the surface to make it hard.
flux: A fusible, nonmetallic material that dissolves and prevents
further formation of metallic oxides, nitrides, or other undesir-
able inclusions within a weld.
forge welding: Joining two metals together by pressure (usu-
ally hammering) with the assistance of heat, but at tempera-
tures below melting.
forging: The hammering and/or pressing of a metal part.
fracture: The rough texture that appears on the surface of bro-
ken metal.
free carbon: The percentage of carbon that is not combined. It
exists in a metal as graphite.
free machining: The property that enables a steel to be ma-
chined more easily, it is a characteristic commonly imparted to
steel by sulfur, which acts as a chip breaker.
fusibility: The relative ease with which a metal melts.
fusion: The melting of metal to the liquid state, permitting two
contacting or neighboring surfaces to partially exchange their
contents, resulting in a thorough blending of the composition
after cooling.
fusion line: The junction between the metal that has been melted
during welding and the unmelted base metal.

243

AppB.pmd 243 10/29/02, 11:28 AM


Tool and Die Making Troubleshooter

fusion weld: A weld created by deposition of molten metal into a


joint formed by two pieces of metal.

G
galling: The abrasion, roughening, and surface welding of a
metal’s surface. Galling develops when two parts are rubbed
together with excessive friction.
galvanizing: The process of hot dipping or electrolytic deposi-
tion of a coating of zinc to finished cold-reduced sheet or fabri-
cated parts made from sheet/strip products.
grain: An individual crystal. Fine-grained steel structures that
result from heat treating are usually preferred to coarse-grained
structures.
graphitic carbon: Free, uncombined carbon in a metallic mate-
rial in the form of flakes or nodules.
gray cast iron: One of the four types of cast iron, it is the most
widely used, one of the least expensive, and has comparatively
less tensile strength than the other three types.
grinding cracks: Tight, shallow surface cracks that initiate per-
pendicular to the direction of grinding on the surface of a
workpiece. Grinding cracks are usually formed by the combi-
nation of residual stresses from heat treatment, excessive heat
from the grinding operation, and the grinding of hard, brittle
materials.

H
H-steels: Alloy steels with tightly controlled elemental composi-
tions specifically for development of hardenability within spe-
cific ranges.
hard crack: See underbead crack.
hardenability: The ability of a steel to harden in depth.
hardening: Any process that increases hardness.
hardness: (1) The resistance of a metal to plastic deformation,
usually by indentation. (2) The degree to which a metal will
resist cutting, abrasion, penetration, bending, and stretching.

244

AppB.pmd 244 10/29/02, 11:28 AM


Appendix B: Glossary

heat-affected zone (HAZ): That area of a metal whose struc-


ture and/or properties have been altered by the heat of a pro-
cess adjacent to it. Welding and electrical discharge machining
(EDM) are two processes that produce heat-affected zones in a
workpiece.
heat check: Surface cracks that develop as a result of thermal
fatigue from alternate and rapid heating and cooling of the
metal’s surface.
heat treating: Any operation that involves heating and cooling
with the express purpose of changing the structures and/or prop-
erties of the workpiece.
hot crack: A crack resulting from stress concentrations in rela-
tively thin weld metal that is the last to solidify. Root cracks
and crater cracks are forms of hot cracking.
hot forming: Any operation that changes the size and/or shape
of the workpiece and is done at temperatures above room tem-
perature. Operations include bending, drawing, forging, press-
ing, and heading.
hot shortness: The characteristic tendency of a material to be-
come brittle at hot working temperatures.
hot shrinkage: A welding condition where the thin weld crater
cools rapidly while the remainder of the bead cools more slowly.
Because metal contracts or shrinks as it cools and shrinkage in
the crater area is restrained by the larger bead, the weld metal
at the center is stressed excessively and may crack.
hot working: Forging and/or rolling of metals at high tempera-
tures.
hydrogen embrittlement: A condition of low ductility in metal
resulting from hydrogen absorption commonly caused by
chrome plating, pickling, or welding.

I
impact strength: The amount of energy a metal can absorb with-
out fracturing under a high-velocity blow.
incipient melting: The beginning of and/or the early stages of
melting.

245

AppB.pmd 245 10/29/02, 11:28 AM


Tool and Die Making Troubleshooter

inclusions: Nonmetallic materials in a solid metallic matrix (for


example, dirt).
India stone: A fine finishing stone used with oil to smooth a die
surface.
induction hardening: The hardening of a ferrous alloy by us-
ing electrical induction to heat it above to the austenitizing tem-
perature prior to quenching.
induction heating: The process of heating by electrical induc-
tion.
inert gas: A gas, such as helium or argon, which does not chemi-
cally combine with other elements.
ingot: A special steel casting for subsequent forging or reducing
into billets or, further, into bars.
ion nitriding: A method of producing nitrided cases using glow-
discharge technology.
isotropic: The quality of having identical properties in all direc-
tions (for example, the absence of a grain direction).

L
laminations: Splits and cracks that may open up during forging,
rolling, machining, or heat treatment.
lap: A surface defect appearing as a seam or deeply impressed
burr, caused by metal folding over during the forging or rolling
operation.
longitudinal grain flow: A grain-flow direction that runs par-
allel to the direction of rolling, drawing, or extrusion.

M
machine straightening: The straightening of material fre-
quently done after rolling and cooling at the steel mill by in-roll
straighteners or machine straighteners, or with special presses.
machining: In general, the removal of surface metal by means
of power-driven metal cutting machinery. More specifically, a
method of conditioning steel by machining away a surface.

246

AppB.pmd 246 10/29/02, 11:28 AM


Appendix B: Glossary

macroetch: Acid etching of a metal surface to accentuate gross


structural details and defects for observation by the unaided eye.
macrostructure: The structure of a metal as revealed by macro-
scopic examination.
magnaflux test: A detailed surface inspection of critical or highly
stressed steel parts. It consists of magnetizing the material and
applying a magnetic powder that adheres along lines of flux
leakage, revealing surface and subsurface nonuniformities.
malleable cast iron: One of four categories of cast iron, it has
superior ductility compared to gray and white cast iron.
martensite: A metallurgical term for the microstructure of a fer-
rous alloy. Martensite is formed by quenching from the
austenitizing temperature at a rate fast enough to form a body-
centered-cubic structure that is highly stressed and hard. When
most tool steels are hardened, their microstructure transforms
to martensite.
medium carbon steels: Carbon steels having carbon contents
in the range of 0.40–0.50% and possibly up to 0.60%. Examples
of steels that fall into this category are 4140, 4340, and 6150
grades.
melting range: The range of temperature at which an alloy melts,
that is, the range between the liquidus and solidus tempera-
tures.
microetch: The cold acid etching usually of a highly polished
and mounted specimen to view its structure at a magnification
of 100× or higher.
microinch: The unit of length equal to one millionth of an inch
(0.000001 in.).
micrometer: Unit of length equal to one millionth of a meter
(0.000040 in.).
microstructure: The structure of metals as revealed with the
aid of a microscope.
mild steel: Carbon steel that has a maximum carbon content of
0.25%.
molecule: The smallest piece of matter into which a compound
can be divided and still maintain all of its characteristics.
ms temperature: The temperature at which martensite begins
to form in an alloy steel on cooling.

247

AppB.pmd 247 10/29/02, 11:28 AM


Tool and Die Making Troubleshooter

N
natural aging: Precipitation hardening of an alloy at room tem-
perature.
nitriding: A process of surface hardening certain types of steels
by heating in ammonia gas at about 930–1,000° F (499–538° C),
with the increase in hardness the result of surface nitride for-
mation. Certain alloying constituents, principally aluminum,
greatly facilitate the hardening reaction. In general, the depth
of the case is less than with carburizing.
nonferrous: Metals or alloys that are relatively free of iron.
normalizing: A thermal treatment used on carbon and alloy steels
to improve structural uniformity, it is achieved by heating to a
suitable temperature above the transformation range and then
quenching in still air.
notch effect: A discontinuity in a part that tends to cause a con-
centration of stresses, thereby weakening the part.
notch toughness: The resistance of a metal to impact forces when
notches or similar irregularities are present.

O
oil hardening: A hardening process in which an alloy steel is
heated to or above the critical range, then quenched in oil.
orange peel: A pebble-grained (rough and pitted) surface that
may develop when too much pressure is applied to a metal dur-
ing polishing.
overheating: Heating a metal to a high temperature that causes
grain coarsening, incipient melting, and/or burning.
oxidation: The addition of oxygen to a compound. The exposure
to atmosphere sometimes results in oxidation of the exposed
surfaces, producing a stain or dislocation, an effect that increases
with temperature.

P
parent metal: The metal to be welded or otherwise worked upon;
also called the base metal.

248

AppB.pmd 248 10/29/02, 11:28 AM


Appendix B: Glossary

passivation: A process that uses nitric acid and/or other chemi-


cals to remove contamination from the surface and/or enhance
corrosion resistance of a metal.
pearlite: A lamellar structure of ferrite and cementite in steel or
cast iron.
peen, peening: Light tapping of the weld bead with the rounded
end of a ball-peen hammer while the bead is hot (approximately
1,200° F [649° C] and still dull red in color), which is sufficient
to cause plastic deformation. Peening minimizes distortion and
improves the toughness and mechanical properties of the weld.
pickling: The process of chemically removing oxides and scale
from the metal surface by the action of water solutions of inor-
ganic acids.
pickup: Metal transfer from tools or from a part to tools during a
forming operation. Sometimes called metal-to-metal welding.
pig iron: Iron produced by a blast furnace and cast into “pigs.”
pit: A sharp depression defect in the surface of a metal.
porosity: Fine holes or pores within a metal.
precipitation hardening: See aging.
precipitation heat treatment: Accelerated aging of a metal at
a temperature higher than room temperature. It is often re-
ferred to as artificial aging.
preheating: Heating of a metal and soaking it at a temperature
below the critical temperature as a preliminary to further ther-
mal or mechanical treatment. Tool steels are generally preheated
prior to heating to the austenitizing temperature to prevent
thermal shock, provide uniformity of heating to the hardening
temperature, and minimize scaling and decarburization.
pyrometer: A temperature-measuring instrument.

Q
quench cracking: The fracture of a metal during quenching from
an elevated temperature.
quench hardening: Hardening a ferrous alloy by quenching it
rapidly enough from a temperature within or above the trans-
formation range to form low-temperature transformation prod-
ucts or martensite.

249

AppB.pmd 249 10/29/02, 11:28 AM


Tool and Die Making Troubleshooter

quenching: Generally, rapid cooling, but more specific terms are


direct quenching, interrupted quenching, selective quenching,
spray quenching, and time quenching.

R
radial crack: A crack originating in the weld fusion zone and
extending into the base metal, usually at right angles to the
line of fusion. This type of crack is due to the high stresses
involved in the cooling of a rigid structure.
recrystallization: A process whereby the distorted grain struc-
ture of cold-worked metals is replaced by a new, strain-free grain
structure during annealing above a specified minimum tem-
perature.
recrystallization temperature: The lowest temperature at
which the distorted grain structure of a cold-worked metal is
replaced by a new, strain-free grain structure during prolonged
annealing.
red hardness: A material’s resistance to softening at elevated
temperatures.
residual stress: Stresses remaining in a material that is free of
external forces or thermal gradients. Residual stresses may be
present from machining, grinding, heat treating, etc.
Rockwell hardness: A standard method for measuring the hard-
ness of metals. The hardness is expressed as a number related
to the depth of penetration of a steel ball or diamond cone
(“braille”) upon application of a designated major weight load.
The penetration is automatically registered as a hardness num-
ber on a dial when the major load is removed. Scales for many
metallic materials range from HRA to HRG, but HRB and HRC
scales are most common for steels and irons.
roll forming: A sheet-metal forming operation involving rolls
that bend, flange, and form.
rolled edge: The edge a universal plate has when rolled by verti-
cal or horizontal rolls. On these plates, edge shearing is not
necessary.
root crack: A weld crack originating in the root bead, which is
usually smaller and of higher carbon content than subsequent

250

AppB.pmd 250 10/29/02, 11:28 AM


Appendix B: Glossary

beads. Cracking is caused by shrinking of the hot metal as it


cools, placing the root bead under tension.

S
scale: The layer of oxide that forms on the surface of steel heated
to high temperatures.
seam defect: A crack-like, longitudinal surface defect on a steel
bar resulting from a discontinuity that is closed during rolling
but not welded.
secondary hardening: The increase in hardness developed by
tempering high-alloy steels at temperatures above those at
which the original and lower hardnesses were achieved.
segregation: The highly concentrated distribution of alloying
elements, and/or impurities, which develops during the solidi-
fication of a steel ingot.
shear: To cut or break by applying forces parallel to the plane of
an object. Also, a cutting operation in which an object is cut
between a moving blade and a fixed edge, or by a pair of moving
blades that may be either flat or curved.
shear failure: Failure that develops when crystals slide past each
other under action of shear stresses.
shear strength: The ability of a steel to resist shear failure.
shear stress: Stress that develops when the crystals of an object
are forced to slide past each other, caused by the application of
forces parallel to its plane.
shot blasting: The cleaning of a metal surface by means of an
air blast, which uses metal shot as an abrasive.
slag: The nonmetallic product that forms on top of molten metal
from the mutual dissolution of flux and nonmetallic impurities
in smelting, steelmaking, refining operations, and welding.
When a bead of weld metal cools, the slag cap on the bead can
be readily chipped away.
slip: Plastic deformation occurring when rows of crystals slide
past each other.
soldering: Joining metals by fusion of filler metals that have rela-
tively low melting points. Most commonly, lead- or tin-base al-
loys are used as soft solders. Hard solders are alloys that have

251

AppB.pmd 251 10/29/02, 11:28 AM


Tool and Die Making Troubleshooter

silver, copper, or nickel bases; use of these alloys with melting


points higher than 800° F (427° C) is generally termed brazing.
solid solution: A solution in which both the solvent and solute
are solid materials at room temperature.
solution annealing: See solution heat treatment.
solution heat treatment: Otherwise known as solution anneal-
ing, solution treatment is the first step in the age-hardening
process. An alloy is heated to a suitable temperature long enough
to allow one or more constituents to enter into solution (to put
precipitating elements into the solid solution). It is then cooled
rapidly to hold the constituent in solution. The metal is soft-
ened rather than hardened and left in a supersaturated, un-
stable state, which may subsequently exhibit age hardening.
This usually applies to some stainless steels, nickel alloys, and
heat-treatable aluminum alloys.
spalling: Cracking and/or flaking of material from a surface.
spheroidizing: A special type of annealing whereby small, round
carbide particles are formed. Tool steels can be machined most
easily in the spheroidized-annealed condition.
spot welding: An electric-resistance welding process in which
the pieces being welded are pressed together between a pair of
water-cooled electrodes and a short-interval electrical current
fuses the small area (spot) at the interface between the pieces.
steel: Refined pig iron or, more specifically, iron combined with
carbon in amounts ranging from a few hundredths of one per-
cent up to 1.7%. All steel also contains varying amounts of man-
ganese and silicon, and when desired, other alloying elements.
strain: A unit of change in size or shape because of applied stress.
strain aging: Aging that occurs when a metal is held for a time
after it is cold worked. Hardness and strength are increased
and ductility is decreased.
strain hardening: The increase of hardness and strength that
occurs from plastically deforming metals below their critical
temperatures.
stress: A deforming force expressed as a load per unit of area.
stress crack: See radial crack.
stress relieving: A process of reducing residual stresses in a metal
by heating to a suitable temperature and holding for a suffi-

252

AppB.pmd 252 10/29/02, 11:28 AM


Appendix B: Glossary

cient time. This treatment may be applied to relieve stresses


induced by casting, quenching, normalizing, machining, cold
working, welding, or EDM.
stringers: Longitudinal inclusions (in the direction of rolling) in
a piece of steel.
superheating: A term that refers to heating a material to too
high a temperature range too rapidly.

T
teeming: Pouring of hot metal into ingot molds.
temper brittleness: A range of tempering temperatures, nor-
mally between 650 and 800° F (343 and 427° C), where steels
become more brittle, that is, they lose toughness or impact
strength upon tempering.
tempering: Heating a hardened steel to a temperature below its
transformation range for the purpose of reducing its hardness
and brittleness.
tensile strength: The point at which a metal will break or crack
when stretched, drawn, or pulled in the longitudinal direction.
thermal conductivity: The ability of a material to conduct or
transmit heat.
thermal expansion: The extent of increase in physical dimen-
sions due to an increase in a material’s temperature.
thermal fatigue: The development of surface cracks in a metal
from expansion and contraction, which takes place from rapid
heating and cooling.
thermal stresses: Stresses in metal resulting from nonuniform
distribution of temperature.
through hardening: A term to describe steels that harden com-
pletely so that the center of a hardened section exhibits the
same hardness as the surface.
toe crack: A crack originating at the junction between the face
of the weld and the base metal. It may be any one of three types:
(1) radial or stress crack, (2) underbead crack extending through
the hardened zone below the fusion line, or (3) crack resulting
from poor fusion between the deposited filler metal and the
base metal.

253

AppB.pmd 253 10/29/02, 11:28 AM


Tool and Die Making Troubleshooter

tool steel: Any high-carbon or alloy steel capable of being suit-


ably tempered for use in the manufacture of tools.
torsion: The act of twisting a material.
toughness: The ability of a steel to resist shock and/or impact
loading without cracking. Toughness is a function of strength
and ductility.

U
ultimate strength: The maximum conventional stress (tensile,
compressive, or shear) that a material can withstand without
cracking.
ultrasonic testing: A nondestructive test employing high-fre-
quency sound to detect internal defects.
underbead crack: A crack in the heat-affected zone of a
weldment just under the weld bead.
unit cell: The smallest part into which an element can be di-
vided.

W
weld metal: A product of fusion, the metal fully melted by the
heat of welding.
white cast iron: One of the four categories of cast iron. It is
comparatively hard and brittle.
work hardening: Same as strain hardening.
workability: The ease with which a metal may be formed.
worm holing: The forming of an irregular cavity as a result of
internal shrinkage during solidification of a cast material. If
the casting is broken open, the cavity can readily be seen by the
naked eye.
wrought: Metals hot worked by hammering or rolling.
wrought iron: A commercial form of iron that is tough, mal-
leable, relatively soft, and contains less than 0.03% carbon.

254

AppB.pmd 254 10/29/02, 11:28 AM


Appendix B: Glossary

Y
yield strength: The point at which an increase in deformation
occurs without an increase in load.

BIBLIOGRAPHY
American Society for Metals. 1948. Metals Handbook, 1948 Edi-
tion. Cleveland, OH: American Society for Metals, pp. 1–16.
ASM International. 1998. Metals Handbook, Desk Edition. Mate-
rials Park, OH: ASM International, pp. 4–63.
Bethlehem Steel Corp. 1980. Modern Steels and Their Properties.
Bethlehem, PA: Bethlehem Steel Corp., pp. 61–73, 191–198.
Iron Age. Undated. Handbook of Terms Commonly Used in the
Steel and Nonferrous Industries. Philadelphia, PA: Chilton.

255

AppB.pmd 255 10/29/02, 11:28 AM


AppB.pmd 256 10/29/02, 11:28 AM
Index

Index

A C
abrasive wear, 50 (Figures 4-7 carbides, 47 (Table 4-3)
and 4-8) carbon
advancements, 208 terms, 45 (Table 4-1)
air quenching, 176 tool steel, 72
aircraft-quality steel, 30 carbonitriding, 207
AISI and SAE steel classifica- carburizing, 206
tion system, 71-72 (Table 5-2) cast-to-shape versus wrought
allotropic, 45 tool steels, 24
alloy segregation, 40 centerline inclusions and
alloying elements, 43, 80 porosity, 19 (Figure 2-8)
annealing, 163, 165 (Table 8-4) Charpy bar, 52-53 (Table 4-4,
argon-oxygen decarburization, Figures 4-9 and 4-10)
15 chemical
atomizing molten metal or analyses of steel, 81 (Table 5-4)
alloys, 20 (Figure 2-9) vapor deposition (CVD), 209,
austenite (retained), 63 213
austenitizing soak cycles, 172 chip, 1, 113-114 (Figures 7-1, 7-2,
(Table 8-8) and 7-3), 124 (Figure 7-13)
chrome plating, 207
color coding for steel types, 70
B (Table 5-1)
banding, 40 compressive strength, 37
breakage or fracture, 39, 124 conduction coefficient, 54
Brinell hardness testing, 196- (Table 4-5)
197 (Figure 9-3) cracked
brittle fracture, 39 die plate, 128 (Figure 7-17)
brittleness, 37 fixtures, 97 (Figure 6-4), 105
built-up edge, 125 (Figure 7-15) (Figure 6-14)

257

index.pmd 257 10/29/02, 11:28 AM


Tool and Die Making Troubleshooter

form die insert, 129 (Figure decarburization, 15, 57-59


7-18) (Figures 4-13 and 4-14), 188
high-speed punch, 171 (Figure deep cryogenic tempering
8-10) (DCT), 210
injection-mold insert, 95 design faults, 92-93
(Figure 6-1) die steel, 85 (Tables 5-7 and 5-8)
injection mold tip, 99 (Figure dimensional changes (shrinkage
6-6) or growth), 189
lathe center, 102 (Figure 6-10) directional properties, 47
milling cutters, 4 (Figure 1-1) discontinuous chip, 113 (Figure
punch after EDM, 133 (Figure 7-1)
7-21) distortion, 167 (Figure 8-8), 183
stamp mark, 103 (Figure 6-11) bending, bowing, or twisting,
tool steel die, 181 (Figure 8-13) 189
tool steel extrusion die, 177 stress, 183-185
(Figure 8-11) volume changes, 185
tool steel punch, 170 (Figure downtime, 5
8-9) ductile fracture, 39
tool-steel punch holder, 99 ductility, 36
(Figure 6-7)
tool-steel swaging die, 104 E
(Figure 6-12) Egyptian iron-smelting furnace,
cracking, 126 8 (Figure 2-1)
craze, 55 elastic limit, 36
during hardening, 190 elasticity (modulus of), 36
fatigue, 123 (Figure 7-12) electric furnace, 10-11 (Figure
fire, 55 2-2)
hardness caused, 169 electrical discharge machining
heat check, 56 (Figure 4-12) (EDM)
mechanical fatigue, 123 effects on tool steels, 130
propagation, 134 (Figure 7-22) electrolysis pitting, 134
quench, 176 wire burning hardened tool
salvaging tools, 127 steel, 131
thermal, 122 (Figure 7-11) electroslag remelting, 16 (Fig-
crater wear, 121 (Figure 7-9) ure 2-6)
crystal structures, 45-46 (Fig- elongation and reduction, 23
ure 4-3) (Figure 2-11)
cyaniding, 207 embrittlement (hydrogen), 39
endurance limit, 37
D
D2 tool steel, 41 (Figure 4-1), F
156 (Figure 8-6), 219 failure modes, 94
Damascus steel, 9 fatigue, 37, 51

258

index.pmd 258 10/29/02, 11:28 AM


Index

ferritic nitrocarburizing, 210 H


file testing, 193 (Figure 9-1)
fine-grained steel, 61 (Figure 4- H13 tool steel, 220-221
16) handling, 3
fire cracking, 55 hardening
flakes, 40 and characteristics, 85 (Table
flame hardening, 207 5-8), 87 (Table 5-10)
flank wear, 120 (Figure 7-7) and cracking, 190
forging, 22 (Figure 2-10) and related stresses, 156
foundation, 204 cycle for D2 tool steels, 156
fracture or breakage, 39, 124 (Figure 8-6)
furnace flame, 207
Egyptian iron-smelting, 8 hardness, 38, 187-188
(Figure 2-1) Brinell, 196-197 (Figure 9-3)
electric, 10-11 (Figure 2-2) carbide, 47 (Table 4-3)
electroslag remelting, 16 caused cracking, 169
(Figure 2-6) conversion, 198-200 (Table 9-1)
loading, 152, 155 (Figure 8-5) non-uniform, 188
of tooling materials, 88
(Table 5-11)
G penetration, 80, 84 (Table 5-6)
galling, 50 relative, 86
gas atmospheres, 159, 206-207 Rockwell, 195 (Figure 9-2)
cautions, 160 surface, 18, 187-188
cyaniding, 207 temper versus, 179 (Figure 8-
endothermic, 160 12)
exothermic, 160 testing, 193, 196-197 (Figure
inert gases, 160 9-3)
nitriding, 206 working, 171 (Table 8-7)
vacuum, 161 heat
grades (steel), 69, 85 (Table 5-7) affected zone, 137 (Figure 7-25)
grain check cracks, 56 (Figure 4-12)
coarsening, 60 checking, 55
direction, 49 (Figures 4-5 and conduction, 54 (Figure 4-11)
4-6), 105 (Figure 6-13) thermal cycles, 149 (Figure 8-2)
grinding treating specification sheet,
cracks (causes of), 126 152
hardened tool steels, 125 treatment, 2, 80, 147, 182
wheel grains act as cutting (Figure 8-14)
tools, 126 (Figure 7-16) Heat Treater’s Lament (A), 148
growth of tool steels, 186 (Table heating guide and tool steel
8-9), 189 properties, 82-83 (Table 5-5)

259

index.pmd 259 10/29/02, 11:28 AM


Tool and Die Making Troubleshooter

high-speed mechanical
steel, 48 (Figure 4-4), 222, 224 etching, 213
steel milling cutters (cracked), fatigue cracking, 123 (Figure
4 (Figure 1-1) 7-12)
tool steel, 74 properties, 35
hot working, 7, 21, 74 metal powder, 20 (Figure 2-9)
hydrogen embrittlement, 39 micro-plasma deposition, 213
mixing shaft, 14 (Figure 2-5)
I modulus of elasticity, 36
incipient melting, 61
inclusions, 13 (Figure 2-4), 19 N
(Figure 2-7), 35 (Figure 3-2) normalizing, 164, 166 (Table 8-5)
ingot solidification, 12 (Figure notch
2-3) effect, 38
insert breakage or fracture, 124 toughness, 38
(Figure 7-14) wear, 120-121 (Figure 7-8)
integral vacuum-furnace
quenching, 176 O
ion, 211-212
O1 tool steel, 226
iron-smelting furnace (Egyptian),
oil quenching, 175
8 (Figure 2-1)
orange peel, 65
ISO 9000 quality system, 27

L P
peening, 141
labor, 3
photomicrograph, 41(Figure 4-1),
lamellar pearlitic microstructure,
131 (Figure 7-19)
116 (Figure 7-4)
physical-vapor deposition (PVD),
208
M pitting, 65, 135 (Figure 7-23)
M2 high-speed steel, 222 plasma
M4 high-speed steel, 224 -assisted (CVD), 213
machinability , 109 source ion implantation (PSII),
material composition, 109, 211
111-112, 115 plastic
of alloy and tool steels, 118 deformation, 122 (Figure 7-10)
(Table 7-1) injection mold cracked tip, 99
prior processing, 110 (Figure 6-6)
ratings, 117 molding tool steel, 75, 86-87
machining, 2, 109, 115 (Tables 5-9 and 5-10)
martensitic stainless steel (420), porosity and inclusions, 13
232 (Figure 2-4), 19 (Figures 2-7
material data, 217 and 2-8), 35 (Figure 3-2)

260

index.pmd 260 10/29/02, 11:28 AM


Index

postheat, 141 say what you do, do what you


powder-metal tool steels, 19 say, 27
preheating, 141, 168 (Table 8-6) scale and surface decarburiza-
profile of a machined surface, tion, 188
119 (Figure 7-6) shear blade, 49 (Figures 4-5 and
profilometer, 118 4-6)
property changes, 3 shock-resisting tool steel, 73
shrink fitting, 142
Q shrinkage, 189
quality, 12, 27-30 (Figure 3-1) single melt versus special
quenching, 173 melting/refining, 18
air, 176 soak, 152, 172 (Table 8-8)
integral vacuum-furnace, 176 spark testing, 42 (Figure 4-2)
oil, 175 special-purpose tool steel, 75
quench cracking, 176 specification sheet, 152 (Figure
rates and stresses, 174 8-4)
vacuum, 163 spheroidized condition, 117
water, 175 (Figure 7-5)
stainless foil wraps, 158
R stainless steel (martensitic,
420), 232
reduction and elongation, 23 stamp-mark cracking, 103
(Figure 2-11) (Figure 6-11)
refining/remelting processes, 14 steel
argon-oxygen decarburization, aircraft quality, 30-31
15 chemical analyses, 81 (Table
electroslag remelting, 16 5-4)
powder-metal tool steels, 19 classification system (SAE
single melt versus special and AISI), 71-72 (Table 5-2)
melting/refining, 18 coarse grained, 60 (Figure 4-15)
vacuum arc remelting, 17 color coding, 70 (Table 5-1)
vacuum degassing, 15 crystal structures, 45-46
vacuum induction melting, 18 (Figure 4-3)
retained austenite, 63 fine grained, 61 (Figure 4-16)
Rockwell hardness, 195 (Figure high speed, 4 (Figure 1-1), 48
9-2) (Figure 4-4), 74, 222, 224
-making and hot working, 7,
S 12 (Figure 2-3)
S5 tool steel, 227-228 mold quality, 31
S7 tool steel, 229 plastics molding, 87 (Table 5-
SAE and AISI steel classifica- 10)
tion system, 71-72 (Table 5-2) stainless (martensitic, 420),
salvaging cracked tools, 127 232

261

index.pmd 261 10/29/02, 11:28 AM


Tool and Die Making Troubleshooter

temperatures, 166 (Table 8-5) cracking, 122 (Figure 7-11)


thermal cycling, 218, 221-222, cycling, 218, 221-222, 224,
224, 226, 228-229, 231-232 226, 228-229, 231-232
strain, 35 diffusion, 209
strength time, 3
comparison of corner sections, tips, 140, 143
106 (Table 6-1) tool
compressive, 37 and die costs, 6 (Figure 1-3)
stress, 35 and-die steel comparator, 76
combinations, 185 (Figure 5-1)
loads, 106 (Figure 6-15) design, 91
mechanically induced, 183 facts of life, 1
-related problems, 64 fixture (cracking), 105 (Figure
relieving, 166 6-14)
thermal, 184 investment, 1
transformation, 184 machining and welding, 109
subzero treatments, 180 material selection, 69
sulfur limits, 45 (Table 4-2) tool steel
surface A2, 217
decarburization and scale, 188 carbon, 72
-finish factors, 117 classifications, 70
hardness, 187-188 cold-working, 72
scale and decarburization, 57 cracks, 97 (Figure 6-4), 102
treatments, 203 (Figure 6-10), 105 (Figure 6-
14), 128-129 (Figures 7-17
T and 7-18)
crimping die, 96 (Figure 6-3)
teamwork, 2, 5 (Figure 1-2) crystal structures, 46 (Figure
temper versus hardness, 179 4-3)
(Figure 8-12) D2, 41 (Figure 4-1), 156
temperature, 166 (Table 8-5) (Figure 8-6), 219
for operations and conditions, decarburization, 15, 57-59
153 (Table 8-1) (Figures 4-13 and 4-14), 188
levels, 154 (Tables 8-2 and 8-3) for stamping dies, 80
references and heat colors, grades, 69
151 H13, 220-221
transition, 39 heating guide, 82-83 (Table 5-
tempering, 2, 179 5)
deep cryogenic (DCT), 210 high-speed, 48 (Figure 4-4),
effects, 64 (Table 4-7) 74, 222, 224
insufficient, 150 (Figure 8-3) hot working, 7, 21, 74
tensile strength, 35 O1, 226
thermal plastic molding, 75, 84, 86-87
conductivity, 53, 55 (Table 4-6) (Tables 5-9 and 5-10)

262

index.pmd 262 10/29/02, 11:28 AM


Index

plate (hardened), 132 (Figure V


7-20)
powder metal, 19 vacuum
properties, 82-83 (Table 5-5) arc remelting, 17
punches, 96 (Figure 6-2), 98 degassing, 15
(Figure 6-5) furnace, 161 (Figure 8-7)
S5, 227-228 heat treatment, 161-163
S7, 229 induction melting, 18
selection, 75, 77-79 (Table 5-3) quenching, 163
shock resisting, 73 volume changes, 185
special purpose, 75
thermal cycling, 218, 221- W
222, 224, 226, 228-229, 231 W1 carbon tool steel, 230-231
toughness, 51 water quenching, 175
W1 carbon, 230-231 wave condition, 62 (Figure 4-17)
wrought versus cast-to-shape, wear, 120
24 and fatigue, 48
tool wear, 48, 120, 203, 206 built-up-edge formation, 124
built-up-edge formation, 124 chipping of the cutting edge,
chipping of the cutting edge, 123
123 crater, 121 (Figure 7-9)
crater, 121 (Figure 7-9) enhancement treatments,
flank, 120 (Figure 7-7) 203, 206
insert breakage or fracture, flank, 120 (Figure 7-7)
124 insert breakage or fracture,
mechanical-fatigue cracking, 124
123 mechanical-fatigue cracking,
milling, 120 123
notch, 120-121 (Figure 7-8) milling, 120
plastic deformation, 122 notch, 120-121 (Figure 7-8)
(Figure 7-10) plastic deformation, 122
thermal cracking, 122 (Figure (Figure 7-10)
7-11) thermal cracking, 122 (Figure
toughness, 38-39, 51 7-11)
transition temperature, 39 welding, 136
troubleshooting, 186 -affected area during welding
cracking during hardening, of cast iron, 136 (Figure 7-24)
190 avoiding pitfalls, 138
dimensional changes, 189 causes of failure, 140
distortion, 189 hardened material, 139
surface decarburization and pickup, 50
scale, 188 preheat, peening, postheat, 141
surface hardness, 187-188 tips, 140

263

index.pmd 263 10/29/02, 11:28 AM


Tool and Die Making Troubleshooter

working hardness, 171 (Table 8-


7)
wrought tool steels versus cast-
to-shape, 24

Y
yield point, 35
yield strength, 35

264

index.pmd 264 10/29/02, 11:28 AM

Вам также может понравиться