0% found this document useful (0 votes)
764 views896 pages

Tool Design - Donaldson

The document is the Special Indian Edition of 'Tool Design' by Cyril Donaldson, George H. LeCain, and V.C. Goold, aimed at undergraduate students and practicing tool designers. It covers various aspects of tool design, including methods, materials, and practices, while incorporating SI metrication and additional solved design problems. The book also includes summaries, questions, and references at the end of each chapter to aid in further study.

Uploaded by

guggillavishal
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
764 views896 pages

Tool Design - Donaldson

The document is the Special Indian Edition of 'Tool Design' by Cyril Donaldson, George H. LeCain, and V.C. Goold, aimed at undergraduate students and practicing tool designers. It covers various aspects of tool design, including methods, materials, and practices, while incorporating SI metrication and additional solved design problems. The book also includes summaries, questions, and references at the end of each chapter to aid in further study.

Uploaded by

guggillavishal
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 896

SPECIAL

INDIAN
EDITION

Tool Design
Fourth Edition
SPECIAL
INDIAN
EDITION

Tool Design
Fourth Edition

Cyril Donaldson
Formerly of Rochester Institute of Technology

George H LeCain
Formerly of Rochester Institute of Technology

V C Goold
Kansas State College of Pittsburg

Joyjeet Ghose
Assistant Professor
Department of Production Engineering
BIT Mesra, Mesra

Tata McGraw Hill Education Private Limited


NEW DELHI

McGraw-Hill Offices
New Delhi New York St Louis San Francisco Auckland Bogotá Caracas
Kuala Lumpur Lisbon London Madrid Mexico City Milan Montreal
San Juan Santiago Singapore Sydney Tokyo Toronto
Tata McGraw Hill
Special Indian Edition 2012
Published by Tata McGraw Hill Education Private Limited,
7 West Patel Nagar, New Delhi 110 008.
Tool Design, 4e (SIE)
Copyright © 1957, 1973, by McGraw-Hill, Inc.
Copyright © 1943 by Rochester Athenacum and
Mechanics institute.
No part of this publication may be reproduced or distributed in any form or by any
means, electronic, mechanical, photocopying, recording, or otherwise or stored in a
database or retrieval system without the prior written permission of the author. The
program listings (if any) may be entered, stored and executed in a computer system,
but they may not be reproduced for publication.
This edition can be exported from India only by the publishers,
Tata McGraw Hill Education Private Limited.
ISBN (13): 978-0-07-015392-9
ISBN (10): 0-07-015392-2
Vice President and Managing Director—MHE: Ajay Shukla
Head—Higher Education Publishing and Marketing: Vibha Mahajan
Publishing Manager—SEM & Tech Voc.: Shalini Jha
Sr Editorial Researcher: Harsha Singh
Copy Editor: Preyoshi Kundu
Sr Production Manager: Satinder S Baveja
Proof Reader: Yukti Sharma
Marketing Manager—Higher Ed.: Vijay Sarathi
Sr. Product Specialist – SEM & Tech Voc.: Tina Jajoriya
Graphic Designer—Cover: Meenu Raghav
General Manager—Production: Rajender P Ghansela
Production Manager: Reji Kumar

Information contained in this work has been obtained by Tata McGraw Hill, from
sources believed to be reliable. However, neither Tata McGraw Hill nor its authors
guarantee the accuracy or completeness of any information published herein, and
neither Tata McGraw Hill nor its authors shall be responsible for any errors, omissions,
or damages arising out of use of this information. This work is published with the
understanding that Tata McGraw Hill and its authors are supplying information but are
not attempting to render engineering or other professional services. If such services
are required, the assistance of an appropriate professional should be sought.

Typeset at Bharati Composers, D-6/159, Sector-VI, Rohini, Delhi 110 085, and printed
at

The McGraw-Hill Companies


CONTENTS

Preface to the Fourth Edition ix


Preface xv
1. TOOL-DESIGN METHODS 1
1.1 Introduction 1
1.2 The Design Procedure 3
1.3 Drafting and Design Techniques in Tooling Drawings 5
Summary 19
Questions 20
References 21
2. TOOLMAKING PRACTISES 22
2.1 Introduction 22
2.2 The Tools of the Toolmaker 23
2.3 Hand Finishing and Polishing 43
2.4 Screws and Dowels 55
2.5 Hole Location 60
2.6 Jig-Boring Practise 75
2.7 Installation of Drill Bushings 90
2.8 Punch and Die Manufacture 96
2.9 Electro Discharge Machining 103
2.10 Electro Discharge Machining for Cavity Applications 113
2.11 Tracer and Duplicating Mills for Cavity Applications 118
2.12 Low-Melting Tooling Materials 125
Summary 130
Questions 130
References 134
3. TOOLING MATERIALS AND HEAT TREATMENT 135
3.1 Introduction 135
3.2 Properties of Materials 135
3.3 Ferrous Tooling Materials 145
3.4 Cast Iron 153
3.5 Mild or Low-Carbon Steel 153
3.6 Nonmetallic Tooling Materials 154
3.7 Nonferrous Tooling materials 155
3.8 Heat Treating 157
3.9 Appearance of Carbon in Steel 161
3.10 Factors Affecting Heat Treating 164
vi Contents

3.11 Heat Treatment and Tool Design 173


Summary 177
Questions 178
Problems 180
References 181
4. DESIGN OF CUTTING TOOLS 182
4.1 Introduction 182
4.2 A Brief History of Metal Cutting 182
4.3 Metal-Cutting Process 183
4.4 Mechanics and Geometry of Chip Formation 187
4.5 General Consideration for Metal Cutting 221
4.6 Metal-Cutting Tools 271
4.7 Single-Point Cutting Tools 272
4.8 Milling Cutters 307
4.9 Drills and Drilling 333
4.10 Types of Drills 337
4.11 Reamers 346
4.12 Reamer Classification 354
4.13 Taps 360
4.14 Tap Classification 363
4.15 The Selection of Carbide Cutting Tools 367
Summary 374
Questions 375
Problems 381
Design Problems 382
References 388
5. GAUGES AND GAUGE DESIGN 390
5.1 Introduction 390
5.2 Fixed Gauges 391
5.3 Gauge Tolerances 403
5.4 The Selection of Material for Gauges 403
5.5 Indicating Gauges 407
5.6 Automatic Gauges 423
Solved Design Problem 425
Summary 428
Questions 429
Problems 430
Design Problems 430
References 433
6. LOCATING AND CLAMPING METHODS 434
6.1 Introduction 434
6.2 The Basic Principles of Location 434
Contents vii

6.3 Locating Methods and Devices 444


6.4 The Basic Principles of Clamping 452
Solved Design Problems 468
Summary 488
Questions 488
Problems 489
Design Problems 489
References 491
7. DESIGN OF DRILL JIGS 492
7.1 Introduction 492
7.2 Definition of a Drill Jig 492
7.3 Types of Drill Jigs 493
7.4 Chip Formation in Drilling 500
7.5 General Considerations in the Design of Drill Jigs 501
7.6 Drill Bushings 507
7.7 Methods of Construction 511
7.8 Drill Jigs and Modern Manufacturing 511
Solved Design Problems 512
Summary 530
Questions 531
Design Problems 532
References 536
8. DESIGN OF FIXTURES 537
8.1 Introduction 537
8.2 Fixtures and Economics 537
8.3 Types of Fixtures 539
8.5 Boring Fixtures 552
8.6 Broaching Fixtures 558
8.7 Lathe Fixtures 561
8.8 Grinding Fixtures 571
Solved Design Problems 573
Summary 583
Questions 584
Design Problems 585
References 590
9. DESIGN OF SHEET-METAL BLANKING AND PIERCING DIES 591
9.1 Introduction 591
9.2 Introduction to Die-Cutting Operations 591
9.3 Power-Press Types 594
9.4 General Press Information 598
9.5 Material-Handling Equipment 598
9.6 Cutting Action in Punch and Die Operations 603
viii Contents

9.7 Die Clearance 605


9.8 Types of Die Construction 618
9.9 Die-Design Fundamentals 626
9.10 Pilots 637
9.11 Strippers and Pressure Pads 639
9.12 Presswork Materials 662
9.13 Strip Layout 663
9.14 Short-Run Tooling for Piercing 666
Summary 671
Questions 671
Problems 674
Design Problems 677
References 679
10. DESIGN OF SHEET–METAL BENDING, FORMING AND 680
DRAWING DIES
10.1 Introduction 680
10.2 Bending Dies 680
10.3 Forming Dies 688
10.4 Drawing Operations 699
10.5 Variables that Affect Metal Flow during Drawing 703
10.6 Determining Blank Size 715
10.7 Drawing Force 716
10.8 Single- and Double-action Draw Dies 716
Solved Design Problem 719
Summary 721
Questions 722
Problems 724
Design Problems 726
References 728
11. USING PLASTICS AS TOOLING MATERIALS 729
11.1 Introduction 729
11.2 Plastic Commonly Used as Tooling Materials 730
11.3 Application of Epoxy Plastic Tools 731
11.4 Construction Methods of Plastic Tooling 734
11.5 Metal-Forming Operations with Urethane Dies 749
11.6 Calculating Forces for Urethane Pressure Pads 760
Solved Examples 762
Summary 763
Questions 764
Design Problems 765
References 768
Contents ix

12. TOOL DESIGN FOR NUMERICALLY CONTROLLED 769


MACHINE TOOLS
12.1 Introduction 769
12.2 The Need for Numerical Control 769
12.3 Basic Explanation of Numerical Control 770
12.4 Numerical Control Systems in Use Today 770
12.5 Fixture Design for Numerically Controlled Machine Tools 775
12.6 Cutting Tools for Numerical Control 791
12.7 Toolholding Methods for Numerical Control 800
12.8 Automatic Tool Changers and Tool Positioners 805
12.9 Tool Presetting 816
Summary 826
Questions 827
Design Problems 828
References 829
13. AUTOMATIC SCREW MACHINE 830
13.1 Introduction 830
13.2 General Explanation of the Brown and Sharpe Machine 830
13.3 Tooling for Automatic Screw Machines 845
Summary 856
Questions 856
References 858
Index 859
PREFACE TO THE
SPECIAL INDIAN EDITION

INTRODUCTION
The book, Tool Design by C. Donaldson, G H LeCain and V C Goold, is universally
accepted for its readability and comprehensiveness.The SIE fourth edition is a value
added SI metrication of this classic book. In addition to serving the undergraduate
students of Mechanical, Production and Industrial Engineering, the present edition
will also be helpful for practising Tool Designers.
In the context of modern tool design, the basic aspects of tool design, devices and
methods remain the same as discussed in the book. However, there has been a para-
digm shift in the tool designing approach, facilitated mainly by the advent of various
computer-aided drafting and designing software programs, augmented with modern
computational facilities. With continuous research worldwide in the field of Tooling and
Tool Design, there has been some significant technological advancement in this field.
In the SIE edition of this book, an attempt has been made to present the book in SI
units and include more solved design problems. A summary of individual chapters
have also been added. The solved design problems added in this edition will prove
to be a valuable material for Tool Design tutorial class exercise or as course work
problems.

Structure of the Book


Chapter 1 defines Tool Design and classifies tooling used in various manufacturing
processes. This chapter presents an overview of the methodology to be adopted in
tool designing practises. A designed tool needs to be manufactured. A toolmaker is
generally entrusted with the responsibility of manufacturing the tool, in accordance
with specified accuracy and tolerances. Chapter 2 familiarises the readers with com-
mon tool making practises.
Tools must be hard, wear-resistant, and have very low surface roughness. Conventional
materials do not cater to these needs; as such tooling requires special materials.
Chapter 3 acquaints the readers with the various available tool materials. Metal cut-
ting tools form an important aspect of tooling applications.
Chapter 4 systematically addresses the geometry, theory, design and related aspects
of cutting tool used in metal cutting. Modern manufacturing requires extensive use of
gauges for shop work. Gages and gauging, therefore is an important aspect of tool
engineering. Chapter 5 delves on the various gauges, gauging process and gauge
design.
Chapter 6 deals with principles, method and devices used for location and clamping
of work. The process of hole making is an integral and difficult part of manufacturing.
The requirement of accurate size hole at an accurate location and in large number
of parts further enhances the complexity of the process. Jigs are suitably applied
xii Preface to the Fourth Edition

for this purpose. Chapter 7 discusses jig design principles, method and devices.
Production jobs often require accurate locating, and holding of the workpiece. This is
achieved using fixtures. Chapter 8 deals with principles, method and devices used
for fixturing.
The process of metal stamping is an integral and difficult part of manufacturing. The
requirement of accurate size blank and accurately pierced sheet metal enhances the
complexity of the process. Chapter 9 deals with methods and devices used for sheet-
metal blanking and piercing and principles of designing sheet-metal blanking and
piercing dies. Blanking and piercing operations are often followed by other related
sheet metal operations like bending, forming and drawing. Chapter 10 deals with
various sheet metal shaping operations like bending, forming and drawing. Modern
manufacturing forced industry to look for new engineering materials. Plastics are
being used as engineering materials, even as tooling materials.
Chapter 11 discusses different types of plastic materials, their applications, and con-
struction of plastic tooling.
Chapter 12 discusses the tooling and fixturing aspects of numerically controlled
machine tools. Automatic screw machine was developed from lathe for economical
manufacture of small screws and bolts. Chapter 13 deals with different aspects of this
specialised machine.
In addition, each chapter ends with Summary, Questions and References to help
students in further study and understanding.

Acknowledgements
I would like to thank all the faculty members of Department of Production Engineering
at BIT Mesra, for their valuable suggestions, in particular, Dr. Vinay Sharma and Dr.
Ritesh Singh. The constant support and continued encouragement of my family and
friends is highly appreciated.
My heartiest thanks to the following reviewers for providing constructive comments:
Dinesh Khanduja National Institute of Technology, Kurukshetra
K S Sangwan Birla Institute of Technology and Science, Pilani
Mohd Zaheer Khan Institute of Technology, Banaras Hindu University, Varanasi
N R De National Institute of Technology, Durgapur
Shailendra Kumar Sardar Vallabhbhai National Institute of Technology, Surat
H Bagchi Sinhgad Engineering College, Pune
A A Shaikh Sardar Vallabhbhai National Institute of Technology, Surat
G L Samuel Indian Institute of Technology, Madras
A Venugopal National Institute of Technology, Warangal
Vijay Desai National Institute of Technology Karnataka, Surathkal
Appu Kuttan National Institute of Technology Karnataka, Surathkal
B K Nanda National Institute of Technology, Rourkela
Preface to the Fourth Edition xiii

I thank all Tata McGraw Hill personnel associated with this project at various stages,
in particular, Mr. Swarnendu Ghosh, Ms. Harsha Singh, Ms. Preyoshi Kundu and
Ms. Yukti Sharma.

Feedback
I solicit your comments, suggestions and criticisms on this adapted version via email
to the publisher or directly to me at the following address: [email protected],
which would be helpful in upgrading the book in the next editions.

Joyjeet Ghose
BIT Mesra, Ranchi
March, 2012
PREFACE

The third edition of “Tool Design” is written as a textbook for beginning courses in tool
design, toolmaking, and related areas of tooling. The content is basic in nature and
covers the major broad general areas of tooling.
Each chapter is designed to give the student an understanding of the fundamentals
of a specific area of tooling. With one or two exceptions, each chapter is indepen-
dent of the others. Chapter 1 serves as an introduction, and thereafter the instructor
may select chapters covering the areas that best suit the needs of his course or pro-
gram. Certain fundamentals of tool design that apply to different areas are sometimes
repeated for this reason.
An attempt has been made to supply the necessary information in each chapter to
enable the student to completely design simple tooling. For example, tables and for-
mulas from various handbooks are included in the chapters where they apply, and
the student should have no problem in their application. It is suggested, however, that
the instructor have handbooks and specialized references available for student use to
supplement the text material. The student should not get in the habit of relying totally
on textbook material, as it is impossible to include all the information he will need,
especially when he begins to design sophisticated tooling. A list of handbooks and
textbooks that apply to the subject matter is found at the end of each chapter.
“Tool Design” is written with the assumption that the student will have a knowledge of
general drafting and manufacturing processes. The units on drafting, therefore, cover
only those aspects which apply to tool design, while the units on toolmaking include
only tools and machines generally used in toolmaking. The student should understand
the basic machining processes of turning, milling, drilling, shaping and grinding.
A student with a knowledge of basic algebra and trigonometry will have no trouble
understanding the information given in this text. For the most part, the mathematics
used is applied rather than theoretical.
Considerable emphasis is placed upon the use of standard parts in the design of tool-
ing throughout the text. Manufacturing firms can no longer afford to pay toolmakers
to individually manufacture common tooling components that can be purchased as
standard items from tooling specialty houses. The goal of the modern tool designer
should be to design a tool made completely of standard parts. The toolmaker’s task
is therefore changed from one of manufacturing to one of precision assembly. The
names of tooling specialty companies are listed at the end of various chapters, and
catalogs may be obtained by writing directly to the company. The instructor is encour-
aged to have available to the student as many different tooling specialty catalogs as
possible to supplement this text.
A considerable amount of the subject matter found in this text will apply to other
areas of study besides tool design. For example, the chapter on the design of cutting
tools is well suited as supplementary material for courses in advanced machine-tool
processes. Chip formation, feeds and speeds, and standard metal-cutting tools are
xvi Preface to the First Edition

probably covered in more detail than can be found in the majority of machine-tool
textbooks. The chapter on tooling for numerically controlled machine tools is also an
excellent supplement for courses in numerical control. Perhaps this is due in part to
the fact that the writer has spent more years in the machine shop than behind the
drawing board. The material presented is therefore written from the viewpoint of the
shop man as well as that of the tool designer.
In conclusion, it should be noted that a major focus of this text is not originality. Rather,
we attempted to collect information on tool design and tooling practices and organize it
in a manner that would make it suitable for use as a basic tool-design text. Information
came from the previous editions of “Tool Design”, handbooks, trade publications,
manufacturer’s literature, textbooks, professional organizations, and from machine
shops and manufacturing firms with which we have had long and close association.
We are grateful to all members of the metalworking manufacturing industries who will-
ingly contributed to the contents of this textbook. Without their help, the publication of
the third edition of “Tool Design” would have been impossible.

Cyril Donaldson
George LeCain
V.C. Goold
TOOL-DESIGN METHODS 1
1.1 INTRODUCTION
Tool design is a specialised phase of tool engineering. Tool-design functions may be
performed by a tool engineer in addition to his other duties in manufacturing, or they
may be performed by a tool-design specialist who devotes his entire working time to
tool design. The size of the company and the type of production generally determine
the extent of specialisation in the tool-design area. A small company may hire only
a tool engineer who is concerned with planning the process of manufacture in addi-
tion to developing tools and machines to produce the product. A larger company may
use a small group of tool engineers who specialise only in designing tools needed to
manufacture the product. A very large company may have a tool-engineering depart-
ment consisting of a tool-design supervisor, several tool designers responsible only
for tool-design layout, and several tool detailers (draftsmen) who finish the drawing.
This text is concerned only with the tool-design aspect of tool engineering, and here-
after all references to that branch of tool engineering will be to “tool design” or “tool
designer”.
The word “tooling” refers to the hardware necessary to produce a particular product.
A considerable amount of tooling is the result of work performed by the tool designer.
Tooling, as viewed by the tool designer, consists of a vast array of cutting devices,
jigs, fixtures, dies, gauges, etc., used in a normal production. The type of production
will determine to a large extent the type of tooling. The most common classification of
types of tooling is as follows:
1. Cutting tools, such as drills, reamers, milling cutters, broaches, and taps.
2. Jigs and fixtures for guiding the tool and holding the workpiece.
3. Gauges and measuring instruments.
4. Sheet-metal pressworking dies for all types of sheet-metal fabrication.
5. Dies for plastic moulding, die casting, permanent moulding, and investment
casting.
6. Forging dies for and cold forging, upsetting, extrusion, and cold finishing.
The tool designer is commonly a specialist in one or perhaps two of the above types
of tooling. For example, a designer of injection moulds for plastics usually has little to
do with the design of metal-cutting tools. On the other hand, a designer of cutting tools
should be knowledgeable in jig and fixture design because of the close relationship
between cutting tools and jigs and fixtures.
The basic task of the tool designer is to provide drawings of a tool or set of tools to
produce the workpiece. He will be provided with a blueprint of the workpiece to be
manufactured, the name and specifications of the machine to produce the workpiece,
and the number of workpieces required. If large numbers of workpieces are needed,
an expensive tool may be justified. If only a few workpieces are needed, the tool must
be inexpensive. In all cases, the tool must be made as economically as possible for
2 Tool Design

the required service. The tool should be easy and safe to operate; it should also look
practical and attractive, but it certainly should not have unnecessary elaborate trim-
mings or needless complexity. The latter point is very important, since even experi-
enced designers sometimes let their enthusiasm for fine mechanisms lead them to
develop excellent tools that are not practical from the standpoint of cost. Of course,
striving to obtain economy may be overdone, and unsatisfactory tools produced. It is
a question of good judgement based on experience. In order to complete his task, the
tool designer may have to produce a complete set of drawings showing (i) an assem-
bly drawing, (ii) one or more subassemblies, if the design is complex, (iii) a detail
drawing of each part, (iv) a complete list of parts needed to make the tool. These are
handed to the toolmakers, whose task is to make the tools.
The tool designer must know the manufacturing procedures. He must be able to
visualise exactly how the workpiece is to be made. He should be competent to judge
the merits of different methods. For example, the tool designer should be able to
determine whether a workpiece A should be made on a shaper or a milling machine
or whether piece B should be a stamping or die casting. The product designer in the
engineering department will also have a share in making decisions relative to the
advantages of a particular stamping to a corresponding die casting. In large concerns,
the tool engineer rather than the tool designer may settle this point, but nevertheless,
the tool designer must understand all this thoroughly to do his job well.
The tool designer must have a knowledge of standards and procedures. The greatest
economy can be effected when standard parts (screws, bushings, handles, claps,
and so on) are used for the manufacture of new parts. Since they are made in large
numbers, standard parts can be manufactured at a lower cost than special jobs.
Furthermore, standard parts can be salvaged from obsolete tools and used again.
A thorough knowledge of procedures is important in modern organisations. This
includes methods used by the plant in manufacturing, in trucking or conveying stock
or parts from one department to another, in inspecting material and products, in draft-
ing, in releasing blueprints and stock lists, and in filling tracings and prints. All this is
largely a matter of experience, but it forms an important part of the assets of a good
tool designer.
In addition, a tool designer must be inventive and original. He must be able to incorpo-
rate his ideas in design layouts. In some of his work, especially as a junior designer,
he may be able to follow older designs, making only slight modifications in order to
meet new requirements. But as he is given more responsibility, the tool designer
generally finds that while his background of experience is invaluable, he is more and
more required to develop original tools. In his work, he should always be ready to try
out new ideas and also ready to abandon them once he sees that they will not work.
Flexibility of mind is important. No designer should be set in his ways. Designs can
always be improved; therefore, suggestions from others should be welcome. It is
sometimes unpleasant to redesign a tool after many hours of drawing, but the tool
designer exists only to produce the best design under the imposed conditions. Time
spent on redrawing will, in the end, cost less than the waste created by an unsatisfac-
tory tool. The beginner must remember that he is not only to rectify mistakes but also
to freely reform structures that may be found unsatisfactory after they were drawn to
scale.
The tool designer must understand how tools perform their function. For this, he needs
a good background in mechanics and mathematics. He should also know the physical
properties of materials used in making tools. These are mostly steel, but since there
Tool-Design Methods 3

are now a large variety of alloy steels, each with its own individual properties, this
subject is a comprehensive one.
A mastery of drafting techniques is as essential to the tool designer as an ability to
read and write. His ideas are valueless unless they can be expressed in a manner
that toolmakers can understand. This means adherence to the standard graphical
language understood by all technical men is essential. But the tool designer has to do
much more than convey ideas. He must convey information in exact terms. All draw-
ings must be clear, complete, exact, and easily understood and must be a genuine
help to the toolmaker. The designer cannot rely on verbal instructions or on the previ-
ous experience of the toolmakers. Instructions like “Make it like the one you made
last time but with two more bushings” lead to trouble and confusion. Such remarks
are impossible in large organisations or where the tools are made by outside jobbing
shops.

1.2 THE DESIGN PROCEDURE


A mistake often made by beginners in tool design is to start the finished design on a
drawing board without prior planning or preparation. The experienced tool designer
does not sit down and systematically assemble a tool on paper unless it is based on
an older design and only slight modifications are necessary in order to make it meet
the requirements. When a new and different design is called for, some type of design
procedure is followed to utilise the designer’s time to best advantage, to prevent mis-
takes, and to bring forth the best and correct design. The design procedure followed
by an experienced tool designer is probably loose and informal because he is able
to take shortcuts he has learned through practise. The following section explains a
rather formal design procedure that works very well for the beginners. As she/he gains
experience, the procedure will become more and more easier, and she/he need not
be conscious of each step.

1.2.1 Statement of the Problem


The first step in the design procedure is to define the problem in a clear and sim-
ple statement of the functional needs. The manufacturing engineer usually provides
the tool designer with the problem. He determines a need for necessary tools and
assigns the job to the tool designer. The tool designer will receive a part print, infor-
mation on why the tool is needed, what must be the capabilities of the tool, the type
of machine the tool must be used on, the number of parts to be produced, and other
pertinent information concerning the par. A considerable amount of discussion may
go on between the manufacturing engineer and the tool designer concerning the tool
in question. Part of the discussion may not be really applicable to the problem and in
fact may cause confusion about what is really applicable to the problem and diverting
from what is really needed. For this reason, a simple statement of the problem is nec-
essary. The problem statement should identify the problem in one or two sentences.
For example, satisfactory statements are “to design a drill jig to hold a support bracket
while drilling three 16 mm diameter holes” or “to design a lathe fixture for holding a
pump housing for in-line drilling and boring of bearing holes.”

1.2.2 The Needs Analysis


The needs analysis, sometimes called the predesign analysis, pinpoints the problem
in terms of functional need. The problem is analysed by asking who, why, how, when,
4 Tool Design

what, and where questions about the functional requirements of the problem state-
ment. All information supplied by the manufacturing engineer is examined, and all
questions that arise are listed on paper to provide a permanent record. The result is a
series of questions pertaining to the requirements of the tool. For example,
1. Will the tool be used by skilled or unskilled operators?
2. How many parts can be held in the tool?
3. What are the hole-location tolerances on the part?
4. How should the part nest in the jig to maintain accuracy in hole location?
5. Will the forces from the clamping device affect hole location?
6. Are pneumatic lines located in the area?
7. What are the measurements of the T-slots on the machine table?
8. Will operating handles of the machine strike the tool?
9. Will the location of the tool prevent removal of cutters for resharpening?
10. Does the operator need more than one size of wrench or loose handle?
11. Are there variations in shape of the parts?
12. Is there any obstruction that may hinder the loading and unloading of the
fixture?
13. What previous operations have been done on the part?
14. If coolant is used, what provisions must be made for coolant passage?
15. What provisions should be made to prevent accumulation of chips?
16. What surface on the workpiece is best suited to locate in relation to the major
reference plan on the machine?
17. Are locating points easily visible to the operator?
18. Will burrs interfere with unloading?
19. Are clamps well out of the way during loading and unloading?
20. If a cutting fluid is used, will knurled knobs make the operator’s fingers sore?
21. Can the tool be made with the available components and facilities?
22. Is the cutting force heavy or light in this operation?
23. Is the cutting force directed towards the solid part of the fixture?
24. Is the work supported directly under the clamping points?
25. Will the clamping force bend the baseplate and cause inaccurate work?
A considerable amount of time should be spent on the needs analysis. Many tool
designers have developed a checklist of needs-analysis questions to make sure no
important points are overlooked. Questions that arise during research and ideation
stage of the design procedure should be added to the list as well.

1.2.3 Research and Ideation (Sketches)


In the research of the design, information and data based on the needs analysis
are gathered. This information will include such items as the dimensions of the part
to be held or produced, the kind of material from which the part is made, the toler-
ances of the part, the dimensions of the machine, the limitations of the machine,
and the amount of tonnage to blank the part. It may be obtained by talking to other
people, taking measurements, making calculations, looking through handbooks and
catalogues, consulting experts, and making experimental mock-ups. Ideas that would
Tool-Design Methods 5

apply to a particular situation should be sketched on paper for future references. It is


also essential to record the source of the information.
While research is being done, many idea sketches should be made. They do not
have to be works of art, but design ideas not recorded in words or sketches many
become lost information. As research and sketching progress, each will affect the
other. Research will lead to new ideas, and the sketches will show the need for more
research. Idea sketches will be combined, eliminated, and reworked and will emerge
as developmental sketches leading to workable design solutions. All sketches should
be kept, although it may seem at the time that a design should be eliminated.

1.2.4 Tentative Design Solutions


The research and sketches should be combined into one or two tentative design solu-
tions, which may consist of rough working drawings showing a side and top view and
perhaps an end view, if needed. They many or may not be to scale, depending upon
the judgement of the tool designer. An isometric or perspective sketch may be made if
desired, although the tool designer is not as concerned with the visual aspects of the
design as the industrial designer is. The tentative design solutions will be evaluated,
the best design selected and reworked, and the final design is decided upon.

1.2.5 The Finished Design


The finished design may not be the actual finished product, for even in the final stages
of drawing changes and additions may be necessary. However, an accurate drawing
must be completed before the toolmaker is able to begin construction. The drawing will
probably consist of a three-view (or more) orthographic drawing, which will be drawn
to scale according to the tool-drawing procedures established by the company.

1.3 DRAFTING AND DESIGN TECHNIQUES IN


TOOLING DRAWINGS
Conventional drafting techniques are followed in tool design with the exception of a
few practises that vary somewhat. The following section explains the differences and
how they are used. No attempt has been made to teach the basics of drafting. It is
assumed that the student has a sound working knowledge of orthographic projection
and is familiar with conventional drafting techniques.
Often tool drawings are used only once, when the tool is constructed. They are
brought back into use only when changes become necessary, such as those caused
by product redesign or changes made to improve tooling performance. They are used
only by highly skilled toolmakers, toolroom personnel, and tooling buyers. For this
reason, many shortcuts can be used in tool drawings, however use of shortcuts would
cause problems on product drawings. Product drawings have a greater circulation
and are used more frequently and usually over a greater period of time, therefore, the
shortcuts used in tool drawings are not permitted on product drawings.

1.3.1 Drafting Practise


The following list of drafting rules generally applies to tool drawings and is intended
as a guide to help maintain uniformity:
6 Tool Design

1. All lines must be dark enough to produce a clear and sharp print. Select a
grade of pencil that will produce lines according to the amount of pressure
applied by the individual. Grade 2H works well for lettering, 4H for object lines,
and 6H for light layout lines.
2. All drawings should be on a standard paper size.
3. All drawing should have a border line from each side of the paper.
4. The material and the title block should be located in the lower right-hand corner
of the drawing.
5. All dimensions should be expressed in inches, with the inch sign omitted.
6. Full-scale drawings should be used whenever possible. Otherwise, use half or
quarter scales.
7. Drawing and dimensioning must help the person who will use the drawing to
make the item in the toolroom. The toolmaker should not have to make calcula-
tions before he can begin producing the tool.
8. Only as many views as necessary to show all required details should be
given.
9. On tool drawings, the part outlines should be shown in red in all views. This
helps the designer while drawing. When black lines on the tool interfere with
red lines of the part, the red lines are drawn as if there were no black lines and
vice versa.
10. Use uppercase engineering lettering throughout the drawing.
11. A name is always assigned to each tool and placed in the title block. The name
usually is the tool name plus the name of the part as noted on the part drawing.
For example, if the name in the title block of a part drawing is “horizontal actuat-
ing rod” the correct title of the drill jig is “drill jig-horizontal actuating rod”.
12. Only critical dimensions, overall dimensions, and location dimensions should
be shown on tool drawings. Dimensions of individual pieces can be indicated
in the bill of materials and need not appear on the drawing.
13. Standard purchased tool components need not be dimensioned. These
include die sets, screws, dowels, springs, knobs, and tooling speciality items.
Dimensions are not necessary because the compenents come in ready-made
form and are identified in the material list by number.
14. Standard purchased tool components that are to be altered by the toolmaker
should have the altered portion dimensioned.
15. Special tooling components that have been standardised by a particular com-
pany do not need dimensions.
16. Dimensions that can be determined by, or calculated from, dimensions on the
part print need not be shown on the tool drawing. Examples would be the cen-
tre of the nest, cutting edges on a punch, die clearance, etc.
17. It is not necessary to dimension the location of screws and dowels in tool
drawings.

1.3.2 Drawing Layout


There are two different methods of preparing tool-design drawings. One is to show
all information, including the details, on one sheet. The tool is shown assembled with
only the necessary views to give pertinent information. Detailed drawings are included
when necessary. The method is generally adopted by companies whose toolmaking
Tool-Design Methods 7

department is such that one toolmaker builds the entire tool or die and does most of
the work on it. This method will be explained in detail in the following sections.
The other method of preparing tool-design drawings is similar to the method of pre-
paring product-design drawings. The assembly is drawn on one sheet, and each com-
ponent is detailed completely on a separate sheet. In this case, the tool is generally
built by several people, each doing one operation on each component. The assembly
may be completed by another person. This allows the company to utilise different
skill levels in the toolroom. This method of drawing also ensures interchangeable
components, which may be a real asset when repairing tools used on continuous
production.

Information blocks These include the title block, stock list and change block. The
title block is generally placed in the lower right-hand corner of the drawing and is
printed on the drawing paper or stamped on with a rubber stamp. It may also be
drawn by a draftsman with the information filled in after the drawing is completed.
The information contained in the title block varies with companies, but generally the
following information is listed:
1. Name of the company
2. Name of the part
3. Name and number of tools
4. Part number
5. Name of the designer
6. Name of the checker
7. Name and number of machine on which tool is to be used
8. Dates
9. Signature of approval
10. Operation name and number
11. General tolerances
12. Sheet number and total number of sheets
13. Scale of drawing
Figure 1.1 shows some typical title blocks.
The stock list is placed directly above the title block and shows the following
information:
1. Detail number (given in detail balloons)
2. Number required
3. Kind of material
4. Rough size of material (may be finished size if no machining is called for; when
special machining is called for, add machining allowances)
5. Name or description of detail (see Table 1.1 for abbreviations)
The stock list should be filled in after the design is completed and all the details
have been assigned a number. Most tool designers begin with detail number 1 at the
bottom of the list and go upward. This practise allows space at the top of the page
for additional details that may be added at a later date because of an engineering
change.
8 Tool Design

TITLE

DRAWN

CHECKED

APPR. DWG. NO.

SHT. OF SCALE

(Company Name, Department and Address)

Drawn By Date Scale


Checked by Date
Ref. Drawings
Machine Name Mach. No.
Part Name Part No.

Tool Name Tool No.

Associated Drawings Drawing No.

APPROVED BY
Name Date (Company Name, Department and Address)

Part Name

Part No.
Operation

Scale
Drawn by Date
No. Sheets A B C D DD Checked by Date
Sheet No.

Fig. 1.1 Typical title blocks

Since one of the major purposes of the stock list is to simplify ordering, an effort
should be made to number and group details according to when they are ordered.
This practise makes the purchasing agent’s task easier, both when ordering and
when checking reviewed merchandise (see Fig. 1.2 for a typical stock list).
The change block is placed in the upper right-hand corner of the drawing; the informa-
tion is entered in a downward direction (toward the stock list). Ample room should be
left for the change block, although it may never be needed. The change block is filled
Tool-Design Methods 9

Table 1.1 Abbreviations commonly used in tool drawings


Abbreviation Meaning Abbreviation Meaning
APPROX Approximately MACH Machine
BP Boiler plate MAN Manual
BUSH Bushing MFG Manufacturer
CI Cast iron MULT Multiple
CS Cast steel NTS Not to scale
CRS Cold-rolled steel OD Outside diameter
C’BORE Counterbore OPP Opposite
C’SINK Countersink PUR Purchase
CHAMF Chamfer PF Press fit
C to C Centre to centre PRC Pierce
DIA Diameter QTR Quarter
DR Drill rod REQ’D Required
DET Detail R Radius
DIM Dimension REL Relief
FAO Finish all over RGH ROUGH
FIG Figure RH Right-hand
FORG Forging SCR Screw
GRD Grind SF Slip fit
GA Gauge S’FACE Spot face
HDN Harden SH Sheet
HEX HD Hexagon head SHLD Shoulder
HSS High-speed steel SOC HD SCR Socket-head screw
HRS Hot-rolled steel SPL Special
HD Head SQ HD Square head
HDLS Headless STD Standard
ID Insider diameter STK Stock
KS Key stock SHT Sheet
MS Machine steel SW Spring wire
MI Malleable iron THD Thread
MAT’L Material TOL Tolerance
MAX Maximum TPR Taper
MIN Minimum TYP Typical
MISC Miscellaneous WELD CONST Welded construction

out only when changes must be made in the design after the drawing is completed.
The original designer may not make the changes, as it is possible for them to be made
several years later.
10 Tool Design

CONTD

10 12 DANLEY DIE SPRING # 920 PUR

9 1 3 DIA × TO SUIT CRS


4
1
8 2 3×3×88 1020

7 1 4×5×6 0–1

6 4 1 DIA × 1 DR
8
5 1 SPRING TO SUIT PUR
1 PUR
4 6 2 DIA × 2 DOWEL
3 1 1 – 20 × 1 SOC HD SCW PUR
4
2 1 31 × 31 × 35 CRS
2 2 8
1 1 DANLY DIE SET—SEE NOTE PUR
DET NO MAT.
DESCRIPTION
NO REQ

TITLE BLOCK

Fig. 1.2 Typical stock list

There are several reasons for changes in the design. An engineering change on the
part may require adaptations. The toolmaker may find mistakes on the part of the tool
designer. Changes may have to be made on the tool when it is brought to production
speed. Or the tool just may not function as the designer had planned.
The following information is shown in the change block:
1. Change symbol
2. Change description
3. Date of change
4. Person making change
5. Person approving change
The alphabet is used as the change symbol. Change symbols are placed as near as
possible to the item being changed. A line should be drawn through the item being
changed without erasing. This provides a record of past figures. In some cases it may
be necessary to erase, as when a new detail replaces an old one. Here the detail
should be erased in all views, and care should be taken not to use the old detail num-
ber. Draw a line through the old detail balloon and add a new one with a new number.
Add the new number to the stock list. Figure 1.3 shows a change block and examples
of methods of denoting change.
Tool-Design Methods 11

Fig. 1.3 Typical change block

Figure 1.4 shows a universal title block often used by contract design and engineer-
ing firms. It may be used in part drawings or tool drawings for virtually any customer.
The revision block is part of the title block and is filled from bottom to top and then
repeated to the left. Space would have to be left for this expansion. Note that the
logotype of the contract design and engineering firm appears in the border rather than
where the proprietary logo normally would appear. This practise allows the contract
designer and engineering firm to place client’s name in the usual logo area.

4 WELDED ASSY. —
1 × 1 × 1 1
4A BRACKET 1020 2 2 4 2 × 2 × 4

RIB 1020 1 × 1×9


4B 2
B.P. 3 × 18 × 36
4C PLATE 4
3 LEVER 1112 1×2×3 MAKES THREE
1 × 1 ×
2 TIP 3140 2 2 1 SEE DET: HON.
1 ANVIL 1020 2×2×4

DET. REQD DESCRIPTION MAT’L. STOCK SIZE REMARKS


AJAX COMPANY

TITLE
SUPPORT BRACKET DRILL JIG
DRAWN J. PRICE 11/17/73
CHECKED JECO 11/18/73
A .100 WAS.125 11/19/73 APPR. WDJ. 11/18/73 DWG. NO.
17-1973
VISIONS DATE LET. DATE SHT. 1 OF 1 SCALE FULL

Fig. 1.4 Universal information block showing title block, stock list, and change block (Jensen
Engineering Co.)

View layout Usually, a minimum of two assembled views of a jig or fixture are nec-
essary to give all pertinent information. A complicated jig or fixture may require three
views of the assembly with sectional views, enlarged views, or detail drawings to
show complicated areas. Detailed drawings should not be used when it is possible to
show dimensions in an assembled, sectional, or enlarged view. Enlarged views aid in
dimensioning a small area. A circle is drawn around the area to be enlarged.
12 Tool Design

Figure 1.5 shows a typical tool-drawing layout for a jig or fixture. Standard procedure
is to show the top views in the upper left-hand corner of the sheet. The side view is
shown in the bottom left-hand corner, with the end view to the right of the side view.
Sectional views are shown in the area between the assembled views and the informa-
tion blocks. However, the position of views may be altered to improve clarity and ease
of dimensioning.

REVISIONS

13 4
12 1
11 1
10 4
1 2 3 4 5 6 7 8 9 10 11 12 13 9 6
8 2
7 1
6 2
5 2
4 2
3 1
2 1
1 1
No. REQD STOCK LIST
TITLE BLOCK

Fig. 1.5 Typical tool-drawing layout for jig or fixture

Figure 1.6 shows a typical tool-drawing layout for a die design. On a die drawing, the
following views are shown:
1. Plan view of die located in the upper left-hand corner of die. It shows a top
view of the die shoe with components mounted to it. The punch shoe has been
removed in this view.
2. Plan view of punch located directly to right of the die shoe. It is shown looking
directly toward the cutting edges of punch or punches. It consists of the punch
shoe with components attached. It appears as though the punch shoe of the
die set had been removed and turned directly over to the right of the die shoe.
3. Cross-sectional view from right to left located directly below the die shoe. The
section is usually taken along the centreline of the die. Both punch and die
shoes are shown, with the die in the closed position. The cross section may be
taken along a jagged lien in order to improve clarity.
Tool-Design Methods 13

Revisions

6
7
Plan view Plan view
of die 8 of punch

13 14 15 16
1 2 3 4 5

9
10
11
12

STOCK LIST
TITLE BLOCK
Cross section from Cross section from
right to left front to back

Fig. 1.6 Typical tool-drawing layout for die design

These views are always shown on die drawings. Complicated dies may require addi-
tional views to improve clarity:
1. Cross section from front to back (die in closed position).
2. Partial cross-sectional view to show details not included in main cross
sections.
3. Side and front view to show details not easily shown in cross-sectional views.
4. Enlarged view with circle drawn around area to be enlarged, used to show
minute details.
Die designs require special dimensions not ordinarily used in other tool drawings.
The shut height of the die is always shown. The shut height of a particular die is the
distance from the top of the punch shoe to the bottom of the die shoe and is shown
on the cross-sectional view of the die. Travel lengths of strippers, pressure pads, and
knockouts should always be dimensioned. All centrelines should be shown. When the
centreline of the part does not coincide with the centreline of the die, the centreline
of the part should be dimensioned in relation to the centreline of the die. Location
dimensions of die details (components) are shown in relation to one or both of the die
centrelines.

Detail balloons Details are identified by numbers located in a straight line and in
numerical order (as near as possible). They are horizontal, vertical or both. Each
14 Tool Design

detail number is enclosed in a “balloon” (circle). The balloon is usually 12 mm in


diameter. A leader is drawn from the balloon to the detail it is to identify. An arrowpoint
attached to the leader should touch the outer edge of the detail. Leaders are generally
curved lines drawn with French or irregular curves. Figure 1.7 shows a typical layout
of detail balloons.

1 12 14 21 30

ALTER

CARB
RC60
HDN

HDN
GRD

FAO
Fig. 1.7 Method of noting heat treatment, special machining, or special treatments on
detail balloons

Detail balloons and leaders should be drawn in only after all dimensioning has been
completed. This practise allows leader lines to be drawn without crossing dimension
lines. Detail 1 should appear in the extreme left balloon and details should be read
consecutively from left to right. Vertical balloons read from top to bottom. Detail 1 is
usually reserved for the die set in die drawings.
Only one balloon is drawn for a group of screws, dowel pins, springs, etc., that are
identical in size. The leader is connected to only one of the identical details in such
a case. The quantity is called out in the stock list. The other identical details are not
identified. Details that are left and right (identical but reversed) should have a separate
detail number.
Every different detail must have its own number except where a standard purchased
assembly is used in the design, for example stock stops, die sets, clamp assemblies,
and gauges. In this case, one number would be given to the entire assembly.
Heat-treating specifications are often shown on the detail balloon leader as shown in
Fig. 1.7. Surface finish, special machining, or special treatments may also be noted
on the balloon leader.

Abbreviations used on tool drawings Abbreviations are used in tool drawings in


a manner similar to that in production drawings. They are especially useful in filling
out the stock list where space is limited. Abbreviations are for the most part stan-
dard; however, there are variations between various companies and industries. For
example, abbreviations used throughout the aircraft industry may be foreign to the
farm-machinery industry.

Detail shortcuts Tool drawings have limited use compared to product drawings, as
explained earlier. The people who use them are familiar with tool-design practises
and can interpret the information given on a drawing without elaborate details. The
following discussion deals with simplifying details in order to save time spent in detail-
ing. Some of the shortcuts verge on the extreme and probably would be used only
within one company. Others would be easily understood between companies.
Tool-Design Methods 15

The main objective in shortcut methods is to save drawing time. Figures 1.8 to 1.12
show the various shortcuts. Figure 1.8 gives various methods of showing dowels,
screws, springs, and cap screws, while Fig. 1.9 shows an assembly using shortcut
methods. Figure 1.10 shows shortcut methods of knurling and cross-hatching. Figure
1.11 covers repeated details. Figure 1.12 shows the use of rights and lefts and the
symmetry concept.
Standard-part tracing templates should not be overlooked as a means of saving time
when preparing tooling drawings. Many manufacturers of tooling specialty items offer
sets of tracing templates covering their entire line of products that will simplify and
save many hours of design and detailing time.
Figure 1.13 shows a typical jig-and-fixture component template printed on high-qual-
ity, translucent white paper so that it can be duplicated right on the blueprint with
any diazo direct-print process equipment. It can be used left or right hand simply by
reversing the template. To duplicate on drawings, a direct-print copy of the desired
component template is made on translucent paper. The duplicated template area is
placed on the drawing and run through diazo direct-print process equipment. The
template shown can also be placed under the drawing for tracing. Additional design
time can be saved by tracing only the heavy outline of each drawing and specifying
the corresponding part number.

Top view Top view


invisible visible Side view
(a)

Top view Side view


(b)

or or

D D

Top view Top view Side view


visible invisible (c)
16 Tool Design

or or

(d)

Fig. 1.8 Methods of showing holes, dowels, screws, and springs in tool drawings (a) Tapped
holes: used hole template (b) Springs (c) Dowel pins (d) Side views: draw with
tolling template (e) Socket-head cap screws

(a) (b)

(c)

Fig. 1.9 (a) Conventional (b) and (c) Shortcut methods of drawing details
Tool-Design Methods 17

(a) (b)

(c) (d)

Fig. 1.10 (a) Conventional method (b) Shortcut for drawing knurling (c) Conventional
method (d) Shortcut for cross-hatching

Fig. 1.11 The elimination of repeated details and dimensions


18 Tool Design

C–L SYM

Fig. 1.12 Use of symmetry shortcut

127
102

89

32

29
35
48
Tool-Design Methods 19

Fig. 1.13 Typical standard jig-and-fixture component tracing template (Jergens, Inc.)

Summary

This chapter defines Tool design and classifies tooling used in various manufacturing
processes. This chapter presents an overview of the methodology to be adopted in
tool designing practises.
• Tool design is the process of designing and developing the tools, methods, and
techniques necessary to improve manufacturing efficiency and productivity.
• Tooling requires manufacturing of complex shape parts of difficult to machine
materials coupled with very high accuracy. Thus, the process of tool design
and its implementation in actual tooling requires large amount of time and
resources. In order to be globally competitive, manufacturers worldwide are in
need for special tooling which can be manufactured within minimum amount of
time and resources, and with the required level of quality.
• Tooling refers to the hardware necessary to produce a particular product.
• Tooling includes cutting devices, dies, jigs, fixtures, gauges, etc.
• A tool designer must be conversant with all available manufacturing processes,
standards and procedures. The tool designer must be clear about the functions
20 Tool Design

that the tool will perform. In addition, a tool designer must be inventive and
original.
• The tool design process includes statement of the problem, the need analysis,
research and ideation, tentative design solution and the finished design.

Questions
1. In your own words, outline the professional requirements of a tool designer.
Mention the courses you have studied that will help you meet these require-
ments. Describe any industrial experience you may have had that will help you
become a good designer.
2. What is the meaning of the word “tooling” as viewed by the professional tool
designer?
3. List and describe the common classifications of tooling types.
4. What is the difference between the toolmaker and the tool designer?
5. What is a design procedure? Why is it necessary?
6. List the five basic steps in a design procedure and briefly explain each step.
7. Why is it necessary to retain all sketches and record where all information was
obtained during the design procedure?
8. Why should all ideas, no matter how vague, be recorded in sketch form during
the design procedure?
9. Why are shortcuts permissible in tool drawings when the same shortcuts would
cause problems in product drawings?
10. Why is the part outline drawn in red on tool drawings?
11. What dimensions are necessary when showing the use of standard purchased
tool components?
12. Why is it not necessary to dimension the location of screws and dowels in most
tool drawings?
13. When listing standard in the stock list, why is it important to include the manu-
facture’s name and stock number?
14. What is meant by the shout height of a die?
15. What basic reference is used when dimensioning locations of die com-
ponents?
16. How many detail balloons are necessary to identify one identical group of
screws?
17. Why are balloon leaders drawn with a French or irregular curve?
18. Why are detail balloons and leaders drawn only after all dimensioning is
completed?
19. Where does the tool designer obtain standard-part tracing templates?
20. Where does the tool designer obtain information on standard parts?
21. List the principal companies manufacturing gauges.
22. Name one company in your area that sells mill supplies, tool steel, and machine
tools.
23. Name two periodicals that deal with production tools and their design.
Tool-Design Methods 21

24. Name one company that supplies standard jig bushings.


25. A number of standard 16 mm. bolts were purchased for a tool being made in
large numbers. Each required a 1.6 mm hole to be drilled in the head. Should
a detail drawing be made? Why?
26. A tool was being designed in one city and manufactured in another. An impor-
tant dimension was omitted from the print sent to the toolmaker. The toolmaker
scaled the print but measured it wrong, and the mistake cost the company
$200. To whose account should it be charged—the toolmaker, the draftsman,
or the checker?

References
Books
• ASTME: “Manufacturing Planning and Estimating Handbook,” F W Wilson (ed.), McGraw-Hill,
New York, 1963.
—: “Tool Engineering: Organization and Operation,” Dearborn, Mich., 1968.
—: “Tool Engineers Handbook,” 2d ed., FW Wilson (ed.), McGraw-Hill, New York, 1959.
• Doyle, L E: “Tool Engineering,” Prentice-Hall, Englewood Cliffs, NJ, 1950.
• Eary, D F, and EA Reed: “Techniques of Pressworking Sheet Metal,” Prentice-Hall, Englewood
Cliffs, N.J., 1958.
• NYSVPAA: “Jig and Fixture Design,” 2 vols., Delmar Publishers, Albany, NY, 1947.
Periodicals
• American Machinist, New York.
• Machine and Tool Bluebook, Wheaton, III.
• Machinery, New York.
• Manufacturing Engineering and Management, Dearborn. Mich.
• Modern Machine Shop, Cincinnati, Ohio.
• Tooling and Production, Cleveland, Ohio.
2 TOOLMAKING PRACTISES

2.1 INTRODUCTION
The tool designer must always keep in mind that the tool he designs must be made
by a toolmaker. He should also remember that the toolmaker generally has to use the
facilities available to him in the toolroom or toolmaking department. It is very important
that the designer be familiar with machine-shop and toolmaking practise and have a
good knowledge of the existing facilities with which the toolmaker has to work. The
tool designer must always remember that it is extremely easy to call out clearances of
less than 25 mm between the punch and dies and extremely difficult to manufacture
them, especially on an odd-shaped contour.
One of the greatest shortcomings of a beginning tool designer is the lack of under-
standing of shop processes and toolmaking practises. He should make every effort to
enroll in courses in shop practise while in school and then spend as much time in the
shop as possible observing after he is on the job. He should become well aquainted
with the toolmakers and ask their advice and opinion. He must always remember that
the expert toolmaker probably knows more about tool design than the average tool
designer. The majority of good tool designers in industry today were first toolmakers.
This is not to say that all good toolmakers would make good tool designers, because
in many cases the toolmaker lacks the ability to communicate and express his ideas
in the form of tool drawings. A technical graduate may make a better tool designer if
he understands machine tools and manufacturing processes. What is important is to
understand that the tool designer and toolmaker must work closely together and that
one is as important as the other.
The tool designer must never appear superior to the toolmaker because he wears
a white shirt and tie. Toolmakers and other shop workers have a natural tendency
to resent the white-collar workers anyway, and every effort on the part of the tool
designer to establish friendly relations will pay dividends in the long run. This is partic-
ularly true for the beginning designer because the toolmaker and machinist can teach
him much that is of value. If the tool designer will remove his jacket, roll up his sleeves,
and enter the shop area not caring that he may get grease on his shirt, he will find that
he will generally be cordially accepted by the shop people, especially if he lets them
know that they are as important as he when it comes to getting the job done.
This chapter is intended to familiarise the beginning tool designer with common tool-
making practises that are generally not mentioned in machine-tool texts. No attempt
will be made to cover common machine-shop or manufacturing practises. There are
excellent texts available for students who wish to read about these subjects. Students
must also understand that there are many other toolmaking techniques that are not
listed in this chapter. A toolmaker develops techniques that are unique to him and his
company and never recorded. The techniques listed here are only a few that have
been practised or observed by the author.
Toolmaking Practises 23

2.2 THE TOOLS OF THE TOOLMAKER


The toolmaker is a generalist by necessity as his work varies from job to job. His tools
are therefore general-purpose tools capable of being applied to a wide variety of jobs.
The standard machine tools and the accessories that make them more versatile are
therefore the tools of the toolmaker. Machine tools that have added features to make
them universal are often called toolmakers’ machines. For example, a small accurate
engine lathe with a wide range of speeds and feeds, along with a complete set of
accessories and attachments to make it more versatile, is often referred to as a tool-
makers’ lathe.
It is apparent that listing all the tools of a toolmaker would be impossible in the few
pages allowed to toolmaking. Standard machine tools and other tools of the general
machinist will therefore not be included. The following discussion will be limited to
special items used by the toolmaker to increase his versatility. A few of the tools dis-
cussed are tools of general machinist, but their importance to the toolmaker makes
their listing essential.

The bench block The bench block shown in Fig. 2.1, is used for holding work when
driving pins, drilling, etc. A V groove across the face accommodates round and odd-
shaped work. The bench block shown is approximately 38 mm high and 76 mm diam-
eter. This tool is generally found on the toolmaker’s bench and is used in the assembly
of small tools and dies.

Fig. 2.1 The toolmaker’s bench block (The L S Starrett Co.)

The toolmaker’s V block and clamp The V block shown in Fig. 2.2, is typically a
toolmaker’s item as it has features to increase its versatility. It can be used on base,
on end, and on either side as the clamp is within the outside width of the block. The
clamp is provided with an adjustable side screw with caps to support the block and
prevent tilting when drilling, grinding, or milling. The groove at the stepped end is at a
right angle to the base and is handy for holding shouldered studs, round pins, etc. A
hole clearance for drilling and removing dowel pins is provided in the block. The block
also has four tapped holes, two in the base and one in each side, so that it can be
attached to a lathe faceplate or held by magnetic chuck.
24 Tool Design

Fig. 2.2 Use of the toolmaker’s V block (The L S Starrett Co.)

The toolmaker’s hammer A small hammer is used for spotting and punching centre
lines and intersections. The hammer shown in Fig. 2.3 contains a magnetising lens
built into the head and eliminates having to look away, as when a separate glass and
hammer are used. High-power magnification makes it easy to spot the punch and
strike without once removing the eyes from the work. The lens is mounted in rubber
for shock resistance.

Fig. 2.3 The toolmaker’s hammer (The L S Starrett Co.)


Toolmaking Practises 25

The toolmaker’s dial test indicator Figure 2.4 shows a typical dial test indicator
that is preferred by toolmakers because of its small size and versatility. The indicator
is mounted so that it can be turned to any angle, and the contact point is so designed
that it can also be swiveled to almost any angle. This type or indicator is very accurate
and should be handled carefully. Because of its accuracy, it is often used to reference
height gauges as well as on machine tools.

Fig. 2.4 The toolmaker’s dial test indicator (The L S Starrett Co.)

The major difference between dial test indicators and regular dial indicators lies in the
method of sensing displacement. Regular indicators sense displacement that is paral-
lel to the axis to the indicator spindle, while test indicators sense displacement that
occurs in a direction perpendicular to the contact point (see Fig. 2.5a). This feature
permits a greater flexibility not thoroughly understood. Although the swivel contact
point can be used at practically any angle with reference to the indicator itself, the line
of motion of the point must be perpendicular to work itself. This is important for accu-
rate readings. Too often a beginner will try to pick up the work as shown in Fig. 2.5b.
When indicator displacement occurs in a direction that is not perpendicular to the
swivel contact point (Fig. 2.5c), the accuracy of the measurement is affected. This
type of measuring discrepancy is commonly known as cosine error. For most work,
this error is negligible. Furthermore, the toolmaker generally uses the dial indicator to
check parallelism and the centreing of circular surfaces, as shown in Fig. 2.6. Here
the toolmaker is interested only in indicator reading variation.
26 Tool Design

There are occasions, however, when the toolmaker is required to determine an accu-
rate indicator reading. When conditions do not permit the swivel-point motion to be
perpendicular to the work surface, corrections for inaccuracy due to setting can be
made by determining angle q (see Fig. 2.5c) and multiplying the indicator reading by
a correction factor.

90°

90°

90°

(a)
(b)

Angle Correct. Angle Correct.


q q factor q factor
80° 0.985 50° 0.766
70° 0.940 40° 0.643
60° 0.866 30° 0.500

(c)

Fig. 2.5 The relationship of the swivel contact point to the workpiece using dial
test indicator (a) Correct (b) Incorrect (c) Correction table for
angle q (multiply indicator reading by correction factor)

Indicator reading shows 0.1652 mm, Angle q = 60°. From


Example 2.1 Fig. 2.5c, the correction factor = 0.866. Therefore, 0.1652 mm ×
0.866 = 0.1431 mm.
Solution: Dial test indicators generally have a switch lever on the body that enables
the toolmaker to reverse the indicator action. In other words, either side of the pickup
may be used, depending upon the position of the switch lever. The bezel can be
rotated to zero the pointer with the dial. An assortment of contacts and attachments
may be obtained for the dial test indicators, including the magnetic base shown in
Fig. 2.4b. A powerful permanent magnet in the base holds firmly to any steel or iron
Toolmaking Practises 27

surface, horizontally, vertically, or upside down, eliminating time spent clamping the
indicator to the machine. A convenient push button turns the magnetic force on or off
for quick one-hand setup and removal of the base.
Figure 2.6 shows the use of the dial test indicator. At a, the indicator is used on a sur-
face grinder to check parallelism of a workpiece held on a magnetic chuck. At b, the
magnetic base is mounted on the lathe cross slide to check the workpiece in a four-
jaw chuck. Centreing a rotary table on the jig borer is shown at c, while the alignment
of a milling-machine vise is shown at d. Centreing a drilled hole for a boring operation

Fig. 2.6 The use of the toolmaker’s dial test indicator (The L S Starrett Co.)
28 Tool Design

on a lathe is shown at e. At f, the indicator is mounted in a vernier height gauge to a


certain reading and then swivels the dial-indicator body until a pickup on the parallel
is shown on the indicator dial (making sure that the contact point is nearly parallel
with the surfaces to be measured). He then rotates the bezel until a zero reading is
shown. The height gauge is adjusted until the indicator contact point is slightly above
the surface to be measured. The contact point is brought against the work surface with
fine adjustment until the indicator again reads zero. A second height-gauge reading is
taken. The distance from the bottom of the binding tool to the top is equal to the differ-
ence in the first and second vernier height-gauge reading.

The diemaker’s squares These instruments, shown in Figs 2.7 and 2.8 are primarily
designed for measuring die clearances. The square shown in Fig. 2.7 contains sliding
blades that can be adjusted at an angle with the beam. The larger knurled thumbscrew
locks the blades at any position, and the smaller one tilts the blades at any angle. To
set the blades at an angle, first release the blade clamp screw; then the blade can be
tilted to the desired angle (determined with a protractor) by turning the small knurled
screw into the beam. The blade can be held in position by tightening the clamp screw.
The offset blade is used when it would be impossible to sight a straight blade.

Fig. 2.7 The diemaker’s square (The L S Starrett Co.)


Toolmaking Practises 29

The square shown in Fig. 2.8 is an improved model in that it is graduated to show the
setting in degrees on the blades. The blades can be set for any angle up to 10° either
side of zero, and the angle is indicated by the line on the pointer.

Fig. 2.8 The imporved diemaker’s square (The L S Starrett Co.)

Transfer tools Figure 2.9 shows the tools commonly used to transfer the location
of existing holes to another workpiece. They are primarily used by the diemaker to
transfer the location of die components to the punch and die shoe of a die set. They
are basically prick punches that centre the punch point in the holes whose location is
to be transferred. Their application to dies is explained later.
The punches shown at a and b are used in the same manner except that the set at b
has an expandable body to fit a wide range of hole sizes. This eliminates the need for
so many punches. The transfer screws at c are used when it is necessary to transfer
the locations of threaded holes. They are especially useful when the threaded holes
are blind.

(a) (b) (c)

Fig. 2.9 Tools used for transferring hole location from one workpiece to another:
(a) Transfer punches (b) Adjustable transfer punches
(c) Transfer screws (The DoAll Company)

Toolmaker’s buttons Toolmaker’s buttons are used to locate holes with positive
accuracy, as explained in detail in Sec. 2.5. Each set consists of four buttons of the
30 Tool Design

same diameter which are fastened to a steel baseplate to protect the screws and
washers used in clamping them to the work (see Fig. 2.54).

Toolmaker’s clamps Figure 2.10 shows a parallel clamp commonly referred to as a


toolmaker’s clamp. such clamps are made of case-hardened steel with the end jaws
chamfered to facilitate clamping under a shoulder or in a recess. They are useful for
holding tool components together during assembly. Their advantage over a C clamp
is that they will not move the workpiece when the screws are tightened. They will also
clamp on nonparallel surfaces to a limited degree.

Fig. 2.10 Toolmaker’s parallel clamps (Brown & Sharpe Mfg. Co.)

Magnetic cylinder squares The magnetic cylinder square shown in Fig. 2.11 is
extremely useful in toolmaking. A magnet in the base of the square securely holds
it to the surface to be checked, as shown in Fig. 2.12. A dial indicator mounted to a
height gauge or surface gauge is used to check height at both ends of the square.
A zero-zero reading indicates that the surface is square is square or perpendicular
to the reference surface. The cylinder squares are available is various sizes. To self-
prove the square, the stem is indicated and rotated 180°. If the square is accurate, the
reading will remain constant.

Fig. 2.11 Magnetic cylinder square (AA Gage Division, US Industries, Inc.)
Toolmaking Practises 31

Fig. 2.12 Use of the magnetic cylinder square as the collet or chuck rotates and approaches
the edge of the work

Edge finders An edge finder shown in Fig. 2.13, is used to pick up the edge of a
workpiece that is to be aligned with the axis of a rotating spindle. In use, it is placed
in 13 mm diameter collet chuck. The head moves sideways as the collet or chuck
rotates and approaches the edge of the work. Location may be made from either flat
or round surfaces or from shoulders. Accuracy is within 13 mm when equipment is in
good shape.

Fig. 2.13 The edge finder (Brown & Sharpe Mfg. Co.)

Sine bars Sine bars measure angles accurately and locate work at a desired angle
to some other surface or line. They are always used in conjunction with some true
surface from which measurements can be taken, preferably a thoroughly clean sur-
face plate.
32 Tool Design

Typical sine bars and sine plates are shown in Fig. 2.14. A sine bar, a sine plate, and
a compound sine plate is shown in Fig. 2.14 a, b, c, respectively. Sine plates contain
tapped holes in the sides, ends and top surface to permit the use of clamps or other
holding devices to hold workpieces securely.

(a) (b)

(c)

Fig. 2.14 Sine bars and sine plates (Brown & Sharpe Mfg. Co.)

The principle of the sine bar is shown in Fig. 2.15. When set in position, the sine bar
forms a right triangle, as shown. A stack of gauge blocks or a planer gauge is used
to construct the side opposite the angle. To set the sine bar to the required angle, it is
necessary to stack gauge blocks up to a height equal to sine bar length multiplied by
the sine of the angle and then set one end of the sine bar on the stack.
A sine bar of particular value to the toolmaker (Figs 2.16 and 2.17) is a small, compact
unit and contains a magnetic base to hold it in place on the reference surface. It is
called the mini sine because of its small size and is extremely useful when making
setups on machine tools. The magnetic base eliminates the problem of holding it in
place and therefore a surface in any plane can be used as a reference surface. The
base contains a V groove which allows the use of a round surface as a reference.
Toolmaking Practises 33

Fig. 2.15 The principle of the sine bar

Fig. 2.16 Methods of setting the mini sine (Kingmann-White, Inc.)


34 Tool Design

Fig. 2.17 Use of the mini sine (Kingmann-White, Inc.)

The mini sine can be set by several methods, as shown in Fig. 2.16. It can be locked
in place after it is set. Figure 2.17 shows various applications of the mini sine. Try to
imagine how some of the applications shown could be made with a standard sine bar
and stack of gauge blocks. Note that a dial indicator is used in combination with the
mini sine.
The sine-bar principle is often built into machine-tool accessories as a means of set-
ting angles accurately. Figure 2.18 shows a grinding-wheel angle dresser constructed
on the sine-bar principle. An accessory of this type enables the toolmaker to dress
grinding wheels to extremely accurate angles.

Fig. 2.18 Grinding-wheel angle dresser constructed on the sine-bar principle


(The DoAll Company)
Toolmaking Practises 35

The locating microscope The locating microscope shown in Fig. 2.19, is used to
pick up edges, contours, irregular shapes, and holes too small for an edge finder or
indicator. The image is not inverted as with standard compound microscopes. This
means the operator sees the work in the same position through the microscope as
without it and that table settings can be made instinctively without the confusion of
transposing reversed or inverted table movements. The reticle reference consists of
a number of concentric circles and two pairs of crossed centre lines that are suitable
for a wide variety of requirements.

Fig. 2.19 The locating microscope (The Moore Special Tool Co.)

Die-handling equipment Die sets are heavy and difficult to handle. The toolmaker
must exert considerable effort in setting and assembling even the smallest die sets.
Many mashed and pinched fingers have resulted from attempting to handle dies with
makeshift carts and pinch bars. Most toolrooms have available portable tables with
adjustable tops to eliminate the need of lifting die sets, as shown in Fig. 2.20. Tables
of this type not only prevent back strain and injuries to fingers and toes but also pre-
vent handling damage to tools and materials. The table at a operates through a hand
crank, worm, worm gear, and screw. It is probably the most positive and safest among
all lift arrangements. It has a capacity of 22 KN and comes in one of six standard
models offered by the manufacturer. The table at b is operated on the hydraulic-jack
principle and lifts loads with little operator effort. Both models are equipped with a floor
brake to prevent unwanted movement.
Figure 2.21 shows die-handling tables for separating, adjusting, and tryout of die sets
prior to placing them in production machines. The table in a contains brackets that
clamp on each side of the die shoe. The die set can then be tilted to the best working
position.
The table in Fig. 2.21b permits die sets and moulds to be separated and refitted and
critical corrections to be made time after time without losing the original alignment.
In use, the die set is positioned on the middle plate, and the top plate is lowered to
contact the punch shoe. The punch shoe is clamped to the top plate and the die shoe
to the middle plate. Turning the lower hand crank rotates the punch shoe to the most
36 Tool Design

convenient working position and may completely invert it. When fitting or corrections
have been made, the punch shoe is rotated to the original position. For tryout, the top
plate is lowered until punch and die engage.

(a) (b)

Fig. 2.20 Portable tables with adjustable tops used for handling die sets and moulds
(a) Hamilton Portelvator (The Hamilton Tool Company)
(b) DoAll elevating table (The DoAll Company)

(a) (b)

Fig. 2.21 Separating tables used to separate die sets and moulds for fitting and
corrections (a) Floor-model die separator (The DoAll Company)
(b) Hamilton Die-part (The Hamilton Tool Company)

Filling machines Filling intricately shaped workpieces to size and shape is an oper-
ation that skilled toolmakers are required to perform. This is a slow and labourious
task when any quantity of material must be removed, and filling machines not only
speed up the operation but increases its accuracy as well.
Toolmaking Practises 37

Filling machines are of two types: the continuous-band filling machine and the recip-
rocating filling machine. The continuous-band machine closely resembles the vertical
metal-cutting band saw with the saw blade replaced with a spring-steel band with
short file segments attached to it (see Fig. 2.22). The segments lock into alignment so
as to eliminate “bump” as the joints pass over the work. A snap joint in the spring-steel
band provides for quick opening and closing of the band loop for internal filling.

Fig. 2.22 The principle of the continuous band file

The reciprocating filling machine resembles a jigsaw and is often called a jig-filling
machine (see Fig. 2.23). In addition to filling, its reciprocating acting makes it suit-
able for sawing and lapping with the insertion of a short metal-cutting saw blade or
abrasive stone.

Fig. 2.23 The reciprocating filing machine (The Oliver Instrument Company)

A band filling machine is shown in Fig. 2.24. Band filling attachments may also be pur-
chased to convert a band saw into a filling machine. The file bands are interchangeable
38 Tool Design

with saw blades. Simply removing the saw guides and replacing them with file guides
makes the band saw a band filling machine.

Fig. 2.24 A continuous-band filing machine (Grob, Inc.)

The major advantage of the band filling machine is that it is always cutting because of
the absence of return strokes. This greatly lengthens file life and helps hold the work
on the table. Angles can be filed by tilting the table.
File bands are available in various width dimensions, shapes and in a variety of num-
ber of teeth per mm and cut designs to suite practically any requirement. Almost
any material can be filed. There are special files for cutting aluminium or bronze, for
cutting boiler-plate steel or for cutting cast iron or alloy tool steel. A bastard-cut file is
predominately used for general filling of steel and other ferrous metals. Permanent
charts are installed on the side of filling machines to guide the selection of files for
specific materials.
When applying the workpiece to the file, any pressure can safely be exerted up to the
point of almost stalling the machine; however, best results are obtained when using
a medium even pressure. A light pressure is best for finish filling. If the teeth of the
file clog, the pressure should be reduced or a band file should be installed with wider
spacing between the teeth. It may occasionally be necessary to use a file card in the
same manner as with hand files.
Band filling speeds vary from 15 m/min for the harder steel alloys to 76 m/min for
softer metals and nonmetals. The chart on the filling machine will give the operator the
correct cutting speed for specific materials.
The advantage of reciprocating filling machines is that intricate shapes can be filed
because of the many shapes of files available (see Fig. 2.25). Internal work is more
easily inspected than on a band file because the overarm can be moved out of posi-
tion with minimum effort and then the workpiece lifted off for gauging or checking. The
table can be titled in two directions for filling of angles and compound angles.
Toolmaking Practises 39

Parallel Machine Files

Round
Half Round
Pillar
(1 Safe Edge)
Oval

Knife

Square
Crochet

Three Square
Pippin
Lozenge
Cant

Fig. 2.25 Parallel machine files used in die-filing machines


(The Nicholson File Company)

Figure 2.26 shows a heavy-duty reciprocating filling machine. A unique feature found
on this machine is an automatic relieving and feeding attachment. This unit applies a
pressure to force the workpiece against the saw or file on the downstroke of the ram
to ensure proper cutting action. Conversely, the pressure is relieved on the upstroke
of the ram to permit accumulated chips to be cleared away. The feeding pressure is
adjustable. This device simply takes much of the manpower away from the operator.
While it does eliminate the conventional need to push the workpiece, it is still neces-
sary for the operator to control the direction of sawing or filling. Note also the small line
for an air blast to keep the workpiece from moving around and are easily adjusted to
the thickness of the workpiece.

Fig. 2.26 Use of a heavy-duty reciprocating filing machine (The Oliver Instrument Co.)
40 Tool Design

The speeds of reciprocating filling machines, given in strokes per minute, is generally
varied by use of a V-belt and stepped-pulley arrangement. The selection of file shapes
and cut and spacing of teeth depends upon the workpiece and workpiece material.
Manufacturers provide this type of information in operators’ manuals that accompany
the machine.

The metal-cutting band saw The contour band sawing machine shown in Fig. 2.27,
is one of the most versatile machine tools found in the toolroom. It is used extensively
for cutting contoured shapes as well as for straight or angled cuts. Clamps or fixtures
are not necessary to hold the work because the cutting force is uniform and towards
the worktable. Workpieces are generally cut to a layout line, as shown in Fig. 2.28,
and sawing radii and irregular contours is possible because of the narrow width of the
band blade.

Fig. 2.27 The contour band sawing machine (Grob, Inc.)

Fig. 2.28 Sawing to a layout line (The DoAll Company)


Toolmaking Practises 41

Contour band sawing machines are available with various worktables, power feed
attachments, coolant facilities, blade welders, polishing attachments, and filling attach-
ments. Figure 2.29 shows the use of the filling attachment.

Fig. 2.29 Use of a filing attachment on a contour band sawing machine


(The DoAll Company)

The simplest machines generally have fixed worktables that are inclinable in two
planes for sawing compound angles. Feeding the workpiece is accomplished by hand
or by a feed mechanism that has a foot-controlled cable arrangement to draw the
workpiece into the saw blade. This relieves the operator from having to push the work-
piece but still enables him to guide it along a layout line (see Fig. 2.30).

Fig. 2.30 Feed attachment on contour band filing machine (The DoAll Company)

The blade-welding attachment is of particular value to the toolmaker because it


enables him to do internal contour sawing on die blocks. In other words, the die open-
ing can be rough-sawed to remove the majority of the material. This is done by drilling
a hole of sufficient size in the die block to accept the blade. The blade is then welded
42 Tool Design

and placed on the machine, where the opening is sawed. The blade must be cut and
rewelded after the sawing operation is completed.

The profile grinder Profile grinders are highly efficient for grinding contours, die
clearances, punches, and many other odd and difficult shapes encountered in tool and
die work. They are available in single- or dual-spindle models. The spindles oscillate
during the grinding operation to provide even grinding-wheel wear and finer finish.
The table tilts forward and back 10° from horizontal, and the graduated quadrant
provides accurate setting for grinding angles. The upper spindle head on dual-spindle
models tilts as much as 10° right or left. A dual-spindle profile grinder is shown in
Fig. 2.31.

Fig. 2.31 The dual-spindle profile grinder (Boyar-Schultz)

The universal punch shaper A hydraulic controlled universal punch shaper is shown
in Fig. 2.32. It works on the hydraulic tracing principle and will produce punch shapes
and contours that are almost impossible to make in any other way. It resembles a
conventional shaper, but the difference lies partly in the design of the table and partly
in the incorporation of the hydraulic system with a copying head. The table design
allows the workpiece to move either in a circular or transverse direction. The work-
piece is usually mounted between centres for rotary shaping, but it can be mounted
in different ways if necessary. The template, which is formed to the required shape
at the ratio of 1:1, is mounted in line with the component at the front of the workhead,
and both move identically, whatever type of shaping is carried out. A stylus rests on
the template, which in turn controls the vertical movement of the table through the
hydraulic system.
Toolmaking Practises 43

Fig. 2.32 A hydraulically controlled punch shaper (The Mercuria Company)

In addition to making punches, the punch shaper is extremely useful in the manufac-
ture of electrodes for use on electrodischarge machines.

2.3 HAND FINISHING AND POLISHING


The final finishing of dies and moulds is often done by hand, depending upon the
type of die or mould and its application. Final finishing may consist of removing less
than 25 mm. from the cutting edge of a die with an abrasive handstone, or it may be
the removal of the small scallops in a die cavity that has been machined with a tracer
mill. Some moulds in the plastics field must be hand-polished to a high luster in order
to impart the necessary surface finish to the product. This section discusses the vari-
ous methods of hand finishing and polishing, along with the equipment that makes
the task easier. It should be noted at this point that hand finishing and polishing is a
labourious and time-consuming job, whatever the method used. Some of the larger
shops that specialise in the production of moulds for plastics hire a crew of women
whose job is to do nothing but finish-polish the mould cavity.

Abrasive sticks Abrasive sticks, often called polishing sticks, stones, or hones, are
manufactured in a multitude of shapes and sizes as well as abrasive materials and
abrasive grain size and hardness. Figure 2.33 shows a few of the shapes readily
available from stock. Abrasive sticks are available in various lengths and every well-
equipped toolmaker will have an assortment of sizes and shapes in his toolbox. The
use of the abrasive stick is shown in Fig. 2.34.

Fig. 2.33 Standard shapes of abrasive sticks (Norton Company, Coated


Abrasive Division)
44 Tool Design

Fig. 2.34 Stoning a die using a round India abrasive stick (The Norton Company,
Coated Abrasive Division)

The two basic classes of abrasive sticks are those containing manufactured abrasives
and those containing natural abrasives. The manufactured abrasive is actually harder
than anything in nature except the diamond. The two manufactured abrasives are alu-
minium oxide and silicon carbide. Stones made of aluminium oxide abrasive are often
referred to as India oilstones. The India stone is unequalled for its long sharpening life,
maintenance of shape, and smoothness of cut. It is recommended for the stoning of
steel cutting edges, where the keeness of the edge is of greatest concern.
Abrasive sticks made of vitrified silicon carbide abrasive are almost as hard as dia-
mond. They are extremely sharp and fast-cutting and will remove metal more rapidly
than any other type of stone. These stones are recommended for operations when the
speed of cutting is more important than the fineness of edge. They appear to be softer
than aluminium oxide stones and therefore do not hold their shape as well.
Abrasive sticks made of natural abrasive are generally referred to as Arkansas stones.
They are cut from deposits of novaculite (a hard, dense, siliceous rock) found in the
Ozark Mountains of Arkansas. Arkansas stones are divided into two grades, hard
and soft, depending upon the density of the stone. Hard Arkansas stone has a very
fine grit and close density. It is generally used for sharpening delicate tools and/or
polishing to a fine finish. It is extremely slow-cutting when compared with other abra-
sive sticks. Soft Arkansas is not as fine-grained and hard as hard Arkansas stone.
Therefore, it cuts faster and is used when a keen, smooth edge is necessary without
the super finish given by the hard Arkansas.
Proper technique in the use of abrasive sticks, as shown in Fig. 2.34, must be devel-
oped through practise and experience. Stoning is a finishing operation, and it takes a
long time to remove even a little material. The beginner soon learns to leave as little
material as possible for stoning.
Toolmaking Practises 45

Abrasive sticks, whether manufactured or natural, soon glaze, and their cutting ability
may be reduced as much as 50 per cent. Water or oil may be used to float residual
from the stone, oil being preferred by most toolmakers. News stones should be soaked
in oil unless they have been presoaked at the time of manufacture. This prevents the
stone from soaking up the oil as it is added during use. In use, a few drops of oil are
placed on the surface of the stone. As the stone is used, the oil spreads and eventu-
ally runs over the edges of the stone, carrying the residue with it. New oil must be
added periodically. Dirty oil should always be wiped off the stone thoroughly as soon
as possible after using it to further prevent glazing. If the stone does become glazed,
it can be cleaned by washing it in gasoline or similar solvent with a stiff brush.

Die files Files are used for the same purpose as abrasive sticks in finishing dies and
moulds, except that their use is limited to tooling materials that have not been heat-
treated. Files remove material faster than abrasive sticks. Ordinary machinist’s files
may be used for rough work where fast removal of material is important. However,
finish filling is generally done with Swiss pattern diesinker’s files and needle files (see
Fig. 2.35).

Needle Files — Round Handle


Round

Half Round

Flat

Crossing

Knife

Square

Three Square

Equaling

Barrette
Joint
(2 Round Edges)
Slitting

Marking

Fig. 2.35 Round-handle needle files used by jewellers, watchmakers, and fine toolmakers
(The Nicholson File Company)

Diesinker’s files are designed for use by toolmakers in dressing and finishing dies of
all kinds. Needle files are used for similar work on a smaller scale. Diesinker’s and
needle files are made in several shapes, as illustrated.
46 Tool Design

Diesinker’s riffler files (Fig. 2.36) are double-ended files curved upward at the ends
and are used for getting into corners, crevices, and holes of intricate dies. They are
primarily used for mould work and are excellent for smoothing the scallops of a mould
produced by a tracer-type diesinking machine.

No. 1
No. 4
No. 5
No. 6
No. 7
No. 8

No. 9
No. 11
No. 12
No. 15
No. 16
No. 18

Fig. 2.36 Diesinker’s riffler files (The Nicholson File Company)

Good filling technique, like that of stoning, is developed through experience. Files will
load and must be kept clean with a file card or brush. The fine cut of Swiss files is par-
ticularly easy to load, and the beginner often makes the mistake of applying too much
pressure, which adds to the problem. Swiss pattern files cut slower than the American
cut machinist’s files, and extra pressure will not make them cut faster.

Power hand-polishing equipment Hand-polishing time can be decreased by using


power hand-grinding equipment, although it will not completely replace hand polish-
ing. Power hand grinders are driven by electric or pneumatic motors. Figure 2.44
shows an electric-powered grinder with the motor built into the handpiece. The speed
of grinders of this type varies from 18,000 to 45,000 rpm, depending upon the preci-
sion and quality of the grinder. Rheostats are available for varying the rpm on indi-
vidual units.
Pneumatic-type hand grinders have a turbine motor built into the handpiece (see
Fig. 2.47 and 2.48). The major advantage of pneumatic grinders varies from 38,000
to over 80,000 rpm. The speed of individual models is varied by an air throttle valve
mounted on the handpiece. Generally speaking, pneumatic grinders are also smaller
and lighter than the equivalent electric model.
Toolmaking Practises 47

A really versatile piece of power hand-polishing equipment is the flexible-shaft machine


shown in Fig. 2.37. The weight of the electric motor is eliminated, and handpieces of
various types and sizes are available. A straight handpiece is shown attached to the
flexible shaft in Fig. 2.37, and a 45° handpiece is shown in Fig. 2.38. Figure 2.39
shows a flexible-shaft machine with a reciprocating or filling handpiece. This hand-
piece converts rotary motion into linear motion and is used for filling, honing, and lap-
ping operations. Figure 2.40 shows the reciprocating handpiece being used for filling.
Riffler files used in the reciprocating handpiece are excellent for finishing the bottom
surfaces of moulds. The speed of stroke can be varied from 0 to 3500 strokes per
minute by use of a rheostat.

Fig. 2.37 Flexible-shaft machine with straight handpiece attached (American


Rotary Tools Company)

Fig. 2.38 A 45° handpiece with various rotating rings used for flat polishing operations
(American Rotary Tools Company)
48 Tool Design

Fig. 2.39 Reciprocating handpiece attached to a flexible shaft (American Rotary


Tools Company)

Fig. 2.40 Filing with a reciprocating handpiece attached to a flexible shaft


(American Rotary Tools Company)

The reciprocating handpiece works well for finishing flat surfaces with abrasive sticks,
as shown in Fig. 2.41. A small piece of abrasive stick is cut from the original, and a
slight indentation is made in its surface with a silicon carbide mounted wheel. A bent
wire with a point on the bent end is mounted into the handpiece. The abrasive stick is
placed in the indentation on the stone. A linear reciprocating motion is thus imparted
to the stone, and yet it is allowed to float free to match the surface to be finished.
Light machine oil applied to the abrasive stick and surface will keep it free-cutting and
prevent loading.
Toolmaking Practises 49

Abrasive stick

Bent wire

Reciprocating
handpiece

Fig. 2.41 Polishing a flat surface with abrasive stick and reciprocating handpiece
for flexible shaft

Rotary files The rotary file is similar to a small milling cutter and is generally made of
high-speed steel or tungsten carbide. Rotary power filling is a fast method of smooth-
ing die cavities after rough machining and is especially well suited for chamfering
corners and forming fillets.
Rotary cutters of this type are classed as files or burrs, depending upon how the teeth
are cut at the time of manufacture. A rotary file has teeth raised with a hammer and
chisel; the teeth of burrs are ground from a hardened blank or milled from a soft blank
and hardened. Ground burrs are more efficient because of the uniformity of teeth and
flutes. Ground burrs can also be resharpened at a fraction of the cost of the original
tool.
Rotary files and burrs are available in many sizes and shapes as shown in Fig. 2.42.
A variety of tooth patterning is also available to meet the requirements of various jobs.
Tooth sizes are classed as coarse, medium, and fine, and tooth geometry includes
positive rake, zero rake, and negative rake.
There appears to be no general agreement on the speed at which a rotary file should
be operated. The toolmaker generally drives a rotary file with a variable-speed motor
and regulates the speed to fit the needs of the particular operations. The speed of
these tools can be guided by the same rules used in other machining operations
with toothed cutters. Speed should vary from 60 to 120 m/min when cutting soft or
50 Tool Design

annealed steel. Tough materials should be cut with a lower speed, while soft, nonfer-
rous materials should be cut much faster. For example, aluminium may be rotary-filed
at 150 to 600 m/min. Generally, speaking, higher speeds will result in better finishes,
but less material will be removed. More material can be removed without clogging
with coarse-tooth cutters. A fine finish can be obtained with a coarse-tooth cutter by
increasing the cutting speed, but an equivalent fine finish can be obtained with fine-
tooth tools operated at slower cutting speed.

Fig. 2.42 Assorted rotary files (Nicholson File Company)

The successful use of a rotary file requires the operator to keep file overhang from the
driving-motor chuck as short as possible. He must also grip near the shank and make
his wrist and arm as rigid as possible. This can be done by resting the wrist of forearm
on the workpiece or bench top when taking a cut. Heavy cuts require the use of both
hands on the driving motor for successful operation.

Mounted wheels A mounted wheel, sometimes called a mounted point, is a small


grinding wheel attached to the end of a small mandrel. It is used in high-speed hand
grinders and is excellent for finishing hardened dies.
Mounted wheels are made in a wide variety of shapes and sizes. They are avail-
able in vitrified (clay) bonds with aluminium oxide or silicon carbide abrasive grains.
Aluminium oxide wheels are best suited for finishing heat-treated die steels, while
silicon carbide wheels are used for cast iron, carbide, and nonferrous materials.
There is a certain amount of risk in using mounted wheels in high-speed hand grind-
ers unless the operator is aware that each mounted wheel has a maximum safe oper-
ating speed. Operations at speeds higher than maximum safe operating speed are
apt to result in failure of the mandrel, either through severe bending or fracture of the
wheel. This is especially true for mounted wheels over 20 mm in diameter. Wheels
smaller than this usually cannot run too fast in the ordinary hand grinder unless it is
an extra-high-speed model.
The maximum safe operating speed for mounted wheels is determined by the size of
the mounted wheel, the size of the mandrel, and the overhang of the mandrel. Wheel
manufacturers publish charts for their products based on these three factors. The
Toolmaking Practises 51

toolmaker would be well advised to check maximum safe operating speeds, especially
when using mounted wheels over 20 mm in diameter in high-speed hand grinders. He
should also chuck as close as possible to the mounted wheel and select mandrels as
large as possible.
In use, the pressure between the wheel and workpiece should not be heavy enough to
cause the mandrel to spring. If the wheel causes burning of the work, it is quite likely
that excessive pressure is being used. A freer-cutting wheel specification may permit
the desired rate of stock removal without excessive pressure.
Mounted wheels may glaze or load in the same manner as large grinding wheels.
Most toolmakers use a hard silicon carbide stick or piece of broken silicon carbide
grinding wheel to dress mounted wheels. This is done by grinding the dressing stick
in the same as grinding the workpiece until the desired result is obtained. Mounted
wheels can also be reshaped in this manner.

Rubberised abrasive points Rubberised abrasive points are mounted points consist-
ing of silicon carbide abrasive held in a bond of premium-grade oil-resistant chemical
rubber. In contrast to hard mounted wheels, the rubber bond has a unique cushioned
action and will cut freely, smoothly, and softly without gouging or digging into the work
surface. It resists clogging or smearing and is ideal for a broad range of applications
when metallic or nonmetallic surfaces must be smoothed and polished without loss
of dimensional tolerance or control. They are excellent for rough polishing moulds, as
shown in Fig. 2.43.

Fig. 2.43 Polishing moulds using rubberised abrasive (Cratex Manufacturing


Company, Inc.)

Rubberised abrasives are available in various standard shapes and sizes, as shown
in Fig. 2.44. Each size and shape is made in four standard grit textures, or composi-
tions, which differ in accordance with the mesh size of the abrasive grain used.
52 Tool Design

Fig. 2.44 Assorted rubberised abrasives (Cartex Manufacturing Company, Inc.)

Care should be exercised when mounting rubberised abrasive points because the
area of greatest stress is at the centre hole. Wheels should be mounted on spindles
or arbors of correct diameter (nominal diameter, +0, – 0.05 mm). The wheel must not
be forced onto the spindle. Proper fit, neither too tight nor too loose, is essential.
The same dressing and truing tools methods are used on rubberised abrasive as on
conventional grinding wheels. A corase vitrified abrasive dressing block is ideal, and
all types of manual or mechanical dressers, as well as diamond truing tools, can also
be used.
As with mounted wheels, the maximum safe operating speeds for rubberised abrasive
points should not be exceeded. Generally speaking, small wheels and points 25 mm
in diameter or less, using mandrels 1 to 5, may be operated at 25,000 rpm. Check
with manufacturers for speeds of larger wheels. The maximum safe speed is based on
mandrel overhang of 12 mm or less. For each additional 6 mm overhang the operating
speed should be decreased by 20 per cent. In light deburring, smoothing, cleaning,
and polishing applications the best speed is usually lower than the maximum safe
operating speed. When applying the wheel to the workpiece or the workpiece to the
wheel, always use a light contact pressure.

Abrasive-cloth rotary tools Abrasive-cloth rotary tools are excellent for finishing
larger dies, such as Kirksite dies, forging dies, forming dies, and deep-draw dies,
moulds, and patterns. They are constructed by folding or rolling abrasive cloth into
various shapes that are mounted on mandrels. The roll type shown in Fig. 2.45 and
the flat disk type shown in Fig. 2.46, are two common shapes. The square pad, shown
in Fig. 2.47, is often used for finishing large sections because the intermittent stroke of
the pad corners avoids the swirling pattern left by disks or wheels. New grit is continu-
ally being exposed as the roll or square pad wear as shown in Fig. 2.45.
Toolmaking Practises 53

Fig. 2.45 Abrasive cloth roll, before and after use, illustrating the uniform wear of a
perfectly balanced roll (Merit Abrasive Products, Inc.)

Fig. 2.46 Using abrasive disks mounted on flexible holders and driven with
pneumatic-powered hand grinders (Merit Abrasive Products, Inc.)

Fig. 2.47 Using an abrasive-cloth resin square pad (Merit Abrasive Products, Inc.)
54 Tool Design

The abrasive-cloth disks with flexible rubber hold-


ers shown in Fig. 2.47 are well adapted to blending
flat surfaces. They also work on contoured sur-
faces because of the ability of the rubber holder to
assume the shape of the contour (see Fig. 2.48)
Abrasive-cloth rotary tools are available in glue-
bond cloth and aluminium oxide abrasive or res-
in-bond cloth and aluminium abrasive or silicon
carbide abrasive. Glue-bond cloth and aluminium
oxide abrasive works best on extreme contours
because it has greater flexibility. It works well on
jobs that require light stock removal. Resin-bond Fig. 2.48 Abrasive-cloth disks
cloth and aluminium oxide abrasive are heavier- with flexible rubber
duty and are used whenever medium to heavy holders are well
stock removal is required. It is probably the best adapted for blending
for blending light welds, edge breaking, fairing, and contours and surfaces
the removal of parting lines and machine marks. (Merit Abrasive Prod-
Resin-bond cloth and silicon carbide abrasive is ucts, Inc.)
recommended for very hard metals, soft aluminium
alloys, rubber, plastics, Haspalloy, and titanium.

Felt bobs Felt bobs are one of the best rotary tools for imparting a high finish to
the workpiece. A felt bob is a shaped dense piece of felt mounted on a mandrel
and impregnated with a polishing abrasive. Standard shapes are shown in Fig. 2.49.
Their major advantages are contourability, resiliency, uniform density, long life, and
economy. They are manufactured in varying degrees of hardness, ranging from extra
soft to flint hard. Each density has its place in polishing and finishing, although four
are most commonly used: soft, medium, hard, and rock hard. Figure 2.50 shows the
use of a felt bob to polish a section of a bending die.

Fig. 2.49 Assorted mandrel-mounted felt bobs (Bacon Felt Company)

Commercial polishing compounds are used for setting up the bobs. When using a
regular grease-base type of compound, (rouge, tripoli, steel-cutting, etc.), it is impor-
tant that the bob be properly coated. This is done by passing the rotating bob lightly
Toolmaking Practises 55

across the compound and then holding it


lightly against the work. Again pass the
compound across the face of the bob a
few times. The bob should now be well
coated and ready to use. As the com-
pound wears off, more should be added
to increase the life of the bobs.
The newer greaseless abrasive com-
pounds are ideal for polishing, deburr-
ing, and touching up. Greaseless abra-
sive compounds, a mixture of grain and
adhesive, are available in bars with
Fig. 2.50 Using felt bobs powered by
grains ranging from very fine to very
an electric hand grinder
coarse. Sizing composition bars are also
(Bacon Felt Company)
available.
To apply the compound, start the tool and after it has reached maximum speed, shut
it off and hold the compound bar against the bob. The frictional heat developed during
slowdown is sufficient to melt the bar and transfer the compound to the bob. After a
few moments, repeat the operation until the desired head is obtained.
For fine microfinishing, diamond compound has no peer, especially when applied
with a felt bob. Manufacturers of injection moulds, wire-drawing dies, and other high-
precision tools requiring high finish depend upon felt bobs and diamond compound.
There is a sequence of applying a diamond compound that should be followed, and it
is well to follow the instruction of the diamond-compound manufacturer.
Final finish polishing can seldom be accomplished in one operation with one grain
size. The general practise is to polish the workpiece with a succession of bobs set
up with different grains sizes. This progression is from a coarse grain to a finer grain,
the number of grains depending upon the workpiece material and the finish required.
Each progression should be reduced by about 30 grain numbers. Many toolmakers
being at a 60 grit and then go to 90, 120, 150, 180, 220, etc. The question of how far
to carry the final grain size depends upon the final finish required.
A good polishing job cannot be hurried. Going from a coarse grit to a fine grit without
using an intermediate grit generally results in more polishing time than if an interme-
diate grit were used, because a great deal more time is required to remove the deep
scratches made by the coarse grain.
Most manufacturers recommend a polishing speed of 1500 to 2200 m/min. Higher
speeds are used on stainless steels and tougher steels. Excessive speed tends to
shorten the life of the bob.

2.4 SCREWS AND DOWELS


The most common method of assembling tool components is by screws and dowels.
Two dowels serve to locate one component in relation to the other and one or more
screws, depending upon the size of the component, firmly hold the component in
place. The general procedure is to drill and tap two screw holes and fasten the com-
ponent in position by gently tightening the two screws. The component can be moved
slightly into final position by light taps with a brass hammer or drift. The toolmaker
56 Tool Design

is able to do this because the screw holes are drilled slightly oversize. Once the
component is in final position, the screws are tightened and component position is
rechecked. Dowel holes are then drilled and reamed while the component is locked
in position. There is generally no attempt made to hold a specific location of dowel
holes because only one tool of a kind is made. The exception to this is tooling built for
continuous production when components subject to wear are standardised for easy
and rapid replacement.
This description oversimplifies the procedure of producing screw and dowel holes. A
general description, like the majority of those found in tool-design texts, does not point
out various factors influencing the production of precision dowel holes. For example,
screw and dowel holes by necessity must be made before heat treatment of the com-
ponent. The component may change dimensionally during heat treatment, and the
dowel holes will not align. The holes themselves may change in diameter and shape.
The toolmaker must be aware of these factors and be prepared to cope with their
effects. The tool designer should also be aware of these factors so that he can be
more sympathetic to the problems of the toolmaker. The following discussion presents
a few of the common practical approaches to screw- and dowel-hole production. It
shall be understood that each individual toolmaker develops his own methods, and
they certainly are not to be condemned as long as he gets the necessary results.

Screw holes and threads Threads generally do not pose a problem with regard to
positional location. Conventional tapping practise serves for most applications for tool-
ing. The tap drill should be drilled through the component whenever possible. Chips
produced by the tap are free to drop through the hole, and the depth of tapping does
not have to be so carefully controlled as for blind holes. Drilling through for tapped
holes also make it easier to transfer the hole location to the joining component.
Component parts are joined by drilling and tapping one component while the other is
clearance-drilled and counterbored so that the socket-head cap screw will be below
the component surface. When a heat-treated component is attached to a component
that is not hardened, as when a die block is attached to a die shoe, it is desirable to
place the threads in the soft component because threads in a part to be heat-treated
sometimes induce cracking when the part is quenched. This is especially true when
the component is made from a water-hard tool steel. Threads do not cool evenly, and
this is just enough to set up stresses that cause a crack through the threads. When
it is necessary to thread components to be hardened, they should be made from an
oil-hard or air-hard tool steel.
The minimum depths of thread (usable thread length) should be at least 1.5 times the
diameter of the screw. When the tool is subjected to heavy shock loads, it is not a
bad idea to increase the thread-engagement depth of two times the screw diameter.
When it is necessary to tap a blind hole, as when there must be a smooth work-
ing surface on the components, the same rules for thread depth should be followed
whenever possible. This sometimes cannot to be done because the height of the
component is restricted. When this occurs, it will be necessary to get as much thread
as possible by using a bottoming tape and to add extra holes to distribute the stress
on the threads evenly.
Good toolmaking practise dictates that all holes to be tapped be relieved by counter-
sinking before tapping. A tap has a tendency to lift a small ring of material from the
holes as it starts. The material that is raised causes a hump on the surface of the
Toolmaking Practises 57

component. This could be removed after tapping, but the removal operation distorts
the first thread somewhat and may cause difficulty in starting the screw. Countersinking
before tapping with a 90° countersink relieves the holes and provides an area in which
the first thread can rise and still be below the component surface. The countersink
should be deep enough to provide a hole surface diameter from 1.6 to 2.4 mm greater
than the major diameter of the thread.
It is generally necessary to remove heat-treat scale and residue from threaded holes
after heat treating unless a controlled-atmosphere furnace is used. A mistake often
made by the novice is to run a tap through the hole after the component has been
heat-treated. This practise dulls the tap, as the component is generally near the same
hardness as that of the tap. Toolmakers sometimes use an old, well-worm tap for this
purpose or a screw with a notch similar to a flute of a tap ground in it. The old tap or
screw is run in and out of the hole until the scale is broken loose.
Often the hole will distort or shrink enough during heat treatment to make the screw
bind in the hole. A small amount of lapping compound placed on the screw and care-
fully worked into the hole with forward and reverse movement will take care of a
specially prepared thread lap to bring the holes to size. The best thread lap is made
of cast iron, as it is porous enough to hold the grains of lapping compound, although
cold-rolled steel will work. Thread a cast-iron rod slightly undersize so that it freely
enters the shrunken threaded holes. Charge the lap with lapping compound and care-
fully rotate it back and forth until the hole is brought to size. The lap may be made
adjustable by splitting the threaded section lengthwise and expanding it by use of a
slightly tapered wedge inserted in the end of the lap. A tapered screw may be substi-
tuted for the tapered wedge when a more elaborate design is desired. Always remove
lapping compound from the tapped hole by washing it with solvent.

Dowels and dowel holes The production of precision dowel holes offers a great chal-
lenge to the toolmaker, as the function of dowels is to provide and maintain accurate
positioning of the component. No real problem is encountered when both components
are soft and will not be heat-treated. All that is necessary is to cap-screw the compo-
nents in place and drill and ream the dowel holes with standard drills and dowel-pin
reamers. Dowel-pin reamers are made undersize the amount necessary to produce
a press fit between the hole and the dowel. The amount of undersize depends upon
the size of the dowel. Dowel-pin reamers were formly made by the toolmaker, but now
they can be purchased commercially either singly or in sets.
Standard dowels are furnished commercially to 5 mm larger than the nominal diameter
with a tolerance of 2.5 mm. If for some reason the toolmaker makes the hole too large,
oversize dowels are commercially available at 2.5 mm and 50 mm over the nominal
diameter. It should be noted that oversize dowels are intended for maintenance work
rather than to cover up the toolmaker’s mistakes.
Dowel holes should always extend through both components whenever possible.
A hardened dowel is next to impossible to remove by conventional methods when
pressed into a blind hole. When it is necessary to use a blind hole in one component,
the blind holes should be lapped enough to remove a tenth or so. This will cause the
dowel to stay in the component with the through hole.
Dowel-hole problems begin with heat-treated components. The holes distort and
shrink enough to cause alignment problems after heat treatment. It becomes neces-
sary for the toolmaker to make the holes undersize before heat treatment and then
58 Tool Design

finish them to size and align them after heat treatment. Dowel holes are finished after
heat treatment by some variation of lapping, honing, or grinding.
For the majority of work, the most practical method of finishing dowel holes in heat-
treated die components is by lapping. Laps are inexpensive, and all that is necessary
in the line of machines is one that will turn the lap. The disadvantage of lapping is that
a great deal of skill is required on the part of the toolmaker. There are certain motions,
touches, and sounds used in the art of lapping which the toolmaker must acquire by
practise and experience.
Various kinds of laps are used in the lapping of holes according to the accuracy
required and the conditions under which the work is done. Typical laps for holes are
shown in Fig. 2.51. Laps made in the toolroom basically consist of a rod of copper,
brass, or cast iron turned to fit the holes to the lapped and split longitudinally for some
distance from the end. The outside diameter of the lap is charged with an abrasive,
and a taper wedge or screw is driven into the end for adjustment as the lap wears.
Larger laps are generally made of cast iron, but the smaller sizes are made of cop-
per or brass because cast iron is too brittle. Many toolmakers consider cast-iron laps
superior to those made of copper or brass; however, the usual hole size for dowels
generally eliminates cast iron as a lap material. Copper seems to be the most com-
mon material used for lapping dowel holes, especially if a large amount of material is
to be removed. A copper lap charged with coarse abrasive cuts quite rapidly, but the
hole should be finished with a different lap charged with a finer abrasive.

Fig. 2.51 Types of shop-made laps used for lapping dowel holes in hardened tool
components

A great deal of time can be saved by using a commercial lap, as shown in Fig. 2.52.
Commercial laps with replaceable copper sleeves are available in various sizes.
Except for the smaller sizes, these laps have an adjustable mandrel on which the
sleeve is expanded to fit the hole. Sizes below 6 mm do not have an adjustment, and
it is necessary to interchange sleeves of various sizes.
Silicon carbide abrasive is used for most dowel-hole lapping and is available in a vari-
ety of grit sizes. It can be obtained as a dry abrasive and mixed with a carrier vehicle
such as kerosene or light machine oil. However, for lapping dowel holes, a commercial
lapping compound already mixed with the carrier vehicle is preferred by most tool-
makers. Commercial lapping compound usually has to be thinned with oil or kerosene
to suit the needs of the individual toolmaker.
Toolmaking Practises 59

Fig. 2.52 Commercial Copper Head laps (The Boyar-Schultz Company)

The lap works on the principle that the abrasive is “charged”, or embedded, in the sur-
face of the soft lap material. The grade or coarseness of the abrasive depends upon
the finish required and the amount that must be removed by lapping. The technique of
charging the lap varies with the individual toolmaker and the results desired. The pro-
cedure for charging a lap with dry abrasive consists of sprinkling the abrasive evenly
on a flat plate and firmly rolling the lap over the powder. Enough pressure should be
applied to embed the abrasive in the surface of the lap. Some toolmakers use a small
brass hammer and gently tap the lap as it is rolled over the abrasive. Excess powder
is removed and the lap is ready to be inserted into the hole.
A charging method used by many toolmakers for commercial laps to be used in dowel
holes is to fill the slot and holes in the copper sleeve with lapping compound. The lap is
then shoved into the hole, and lightweight machine oil or kerosene is liberally squirted
around the laps to float the abrasive. When the lap is turned, the abrasive will carry
around and gradually work out evenly to the ends of the hole. The liberal application
of oil floats loose abrasive away from hole openings, and the lap should be expanded
occasionally as lapping progresses without adding new abrasive. Excess abrasive or
new abrasive added to the lap at the hole opening will produce a bell-mouthed hole.
This is why a lap should always be shoved into the hole completely before it is rotated.
Many beginners make the mistake of sprinkling abrasive on the lap and shoving it into
the work while it is turning. This practise will invariably produce a bell-mouthed hole.
Laps for dowel holes are generally powered by a sensitive drill press. There is often
controversy over how fast to rotate a lap. Handbooks recommend speed ranging from
100 to 250 m/min, but a speed this fast is often impossible to obtain when lapping
small holes. Most toolmakers try to rotate the lap at as fast a speed as possible and
yet control the lapping operation, which usually calls for a considerable amount of trial
and error.
The lap should be kept tight enough in the hole to create heat during the operation.
The lap should always fill the hole. This is simplified when using an adjustable lap, but
a solid or commercial sleeve lap in the smaller sizes without adjustment complicates
60 Tool Design

matters. A series of sleeves in small size increments must be used to complete the
lapping operation. Care must be taken to use each size only until the next larger
sleeve in the series will enter the hole.
The lap selected should always be longer than the hole depth in order to straighten
the hole as it is lapped. Long holes of small diameter are difficult to lap straight, and
it may be necessary to relieve the dowel holes to reduce the hole length. As the lap
revolves, it should be slowly drawn back and forth from end to end, and under no cir-
cumstances should this oscillation cease while the lap is rotating.
Components for heat-treated tools are screwed in place before heat treatment, as
described earlier. Dowel holes are then drilled and reamed undersize with compo-
nents in place. How much undersize is required, depends upon how much the par-
ticular steel will distort during heat treatment. This may run from 13 to 130 mm, and
the toolmaker will have to rely on his past experience when determining this. The
components are then heat-treated and perhaps surface-ground, depending upon the
application of the tool. The components are then again bolted together and the dowel
holes lined up as close as possible. Scale and sharp edges can be removed from the
hole by a soft-steel bar stock slightly smaller that the hole with a longitudinal groove or
split in which the edge of a piece of emery cloth is inserted. The cloth is wound around
the bar until it fills the hole in the work. The bar is rotated in the hole until all scale and
rough edges are removed. The hole is now ready for lapping, as previously described,
until it reaches the diameter for the desired dowel. Care must be taken to remove all
abrasive from the hole before checking and fitting the dowel in the hole.
Using a jig grinder to produce dowel holes in hardened components is best. The dowel
hole can be ground with locational accuracy if required, as well as being perfectly
round. In addition, perpendicularity of the hole to the base surface is assured. The
hole can be produced with little effort compared to lapping. When tools require inter-
changeability or replacement of components, the jig grinder becomes a necessity for
producing dowel holes.

2.5 HOLE LOCATION


The problem of locating holes accurately has plagued the toolmaker since the begin-
ning of the industrial revolution. Left to his own devices, the toolmaker has come
up with some ingenious ideas and methods which may or may not be successfully
applied by other toolmakers. The degree of accuracy has often depended upon the
individual skill of the toolmaker, especially when he utilises only conventional machine
tools and improvised tooling methods.
Hole location and production are divided into three basic operations: (i) Establishing
the correct position of the finished hole, (ii) Machining the hole, and (iii) Checking the
hole for accuracy. Establishing the location and machining the hole are probably the
most difficult of the three. There is a big difference between measuring the accuracy
of existing holes and machining them in the desired location to a predetermined size
because of the many variables the toolmaker has to work with during the locating and
machining operation. Examples are the transfer of measurements that result in cumu-
lative errors, inaccuracy of machine tools, temperature differential between the point
of cutting and the rest of the workpiece, stress in workpiece material, manual skill of
the toolmaker, and deflection of the workpiece, cutting tool, and machine tool.
Toolmaking Practises 61

The following discussion describes various methods of locating and machining holes
without reference to their applications. The methods range from simple bench oper-
ation to sophisticated jig-boring machines. The methods used in actual application
depend upon the accuracy required, the skill of the individual toolmaker, the machine
tools and measuring tools available, and the number of holes to be produced.

Layout Layout means scribing lines on the workpiece at right angles to each other.
The intersection of the lines indicates the desired position of the finished hole. Layout
in its simplest form is accomplished by use of a straightedge, combination square, and
hand scriber. This is the least accurate of the layout methods because of the necessity
of aligning the scriber and workpiece with the scale graduations and the inconsistency
of scriber angle against the scale. This method is generally used by the toolmaker for
rough layout for reference when doing precision location on a machine.
Other layout methods listed in order of increasing accuracy are use of the surface
gauge, height gauge, planer gauge, and gauge blocks in combination with a lapped
scriber. These methods are well described in detail in most basic machine-shop texts,
and time will not be devoted to them here except for some general comments.
The surface gauge is only a little better than a hand scriber and straightedge, as it
eliminates the variation in scriber angle and is used in reference to an established
plane (generally a surface plate). It is still necessary to set the scriber with refer-
ence to a scale, and transfer, but transfer of measurements from the scale to the
surface-gauge scriber point often results in a slight error. The height gauge eliminates
this shortcoming, as it is set directly without transfer, but it is slower because of the
necessity of reading the vernier and then
doing arithmetic to three decimal places.
Using a planer gauge and scriber set with
a micrometer is slightly more accurate but
less handy. Precision gauge blocks and a
lapped scriber (Fig. 2.53) are the ultimate
in layout accuracy but of little use because
the toolmaker cannot pick up the intersec-
tion of the layout lines with the cutting tool
to the accuracy of the gauge blocks. Many
toolmakers highlight the layout-line inter-
section with a prick-punch mark, which
further reduces the accuracy of any layout
method. The toolmaker or machinist who
can consistently hit the exact intersection Fig. 2.53 The use of a special lapped
of layout lines made with a height gauge scriber and gauge blocks for
is a rare individual. What he generally layout (The Moore Special
accomplished is to obscure the intersec- Tool Company)
tion to the extent that he is never sure
where the true intersection lies.

Toolmaker’s buttons The use of toolmaker’s buttons to establish the desired loca-
tion of a finished hole in the workpiece has been used in the past to overcome the
limitations of layout methods. Toolmaker’s buttons are small cylinders of hardened
steel which are accurately ground and lapped to size. The ends are ground square
with the sides to assume perfect alignment with the work. Each set consists of four
62 Tool Design

buttons of the same diameter which are fastened to a steel baseplate to protect the
screws and washers used in clamping them to the work. Buttons are available in vari-
ous diametric sizes.
Preliminary preparations for use of tool-
maker’s buttons is to lay out hole positions
by one of the layout methods. The approx-
imate hole positions are then drilled and
tapped to the size of the screws used to
attach the buttons. The button is fastened
to the workpiece and roughly located by
means of scribed lines. The holes in the
buttons are larger than the screws, to
permit sideways adjustment after the pre-
liminary fastening to the workpiece. The
screw is snug, but still loose enough to
permit movement, so that it can be moved
to the desired position. The button is Fig. 2.54 The use of a height gauge and
generally moved by use of a small brass dial indicator to check the posi-
drift tapped lightly with a small hammer. tion of toolmaker’s buttons (The
Various methods used to check the loca- Moore Special Tool Company)
tion of the buttons in relation to each other
on the reference edges of the workpiece include height gauges, micrometres, sine
bars, adjustable parallels, gauge blocks, etc. Figure 2.54 shows the use of a height
gauge and dial indicator to check button positions.
Locating toolmaker’s buttons is a painstaking and time-consuming operation since the
button must be aligned in two directions. A tap in one direction will invariably displace
the button in the other. When the button is finally located, the screw must be tightened,
which may again shift the button.
When all buttons are finally located and tightened, the workpiece is mounted on a
machine tool and one button is located directly in line with the spindle axis by moving
the workpiece. Final location is accomplished by indicating the button with a dial indi-
cator. The button is then removed and the hole is rough-drilled. A single-point boring
tool is used for final truing and sizing of the hole. The workpiece is loosened and the
next button is located and the process repeated until all holes are completed.
The machine tools most commonly used to produce holes located by toolmaker’s
buttons are the vertical milling machine and the engine lathe. In the early days, often
the only machine available for precision machining was the engine lathe. The work-
piece was located on the faceplate, and holes were produced by drilling and boring.
Toolmaker’s buttons were the only practical way of locating holes in relation to each
other because the workpiece had to be unclamped and moved for each successive
hole. For all practical purposes, toolmaker’s buttons are of little use today in the mod-
ern toolroom because machines have been developed for much faster precision locat-
ing and boring of holes.

Transfer Toolmakers often use transfer methods for establishing the location of
holes, especially to match the positions of existing holes in a tool component with cor-
responding holes in another component. Generally, the alignment of holes in the com-
Toolmaking Practises 63

ponents is more important than their location to dimension. Often the exact location of
the original hole is not determined.
The transfer screws and punches are the most common means of transferring hole
locations. These special tools have a small prick-punch point in their centres and
are guided by the holes in the existing tool component. This type of tool is espe-
cially well suited for rough location of punches with respect to die blocks within a
die set. The die block is mounted in place by use of screws and dowels to the dies
shoe of the die set. The die set is inverted and placed on parallels, as shown in
Fig. 2.55.
The punch is placed on the punch shoe, and the assembly is closed to the point where
the punch enters the die block. Parallels should be used to prevent the assembly from
closing too far. Shim stock equal to the die clearance is used to centre the punch prop-
erly in the die block. Rubber bands or soft copper wire are often used to aid in holding
the shim stock in place. The punch is now ready for spotting.
The method of transfer or spotting will depend upon the location of the holes. A trans-
fer punch can be used as shown in Fig. 2.55a when the holes are circumscribed by
the opening in the die shoe. Transfer screws or buttons are used as shown in Fig.
2.55b when the holes cannot be reached through the die opening. When transfer
punches are used, the punch is tapped lightly to indicate the hole location. When
transfer screws are used, a brass drift is placed on the punch as near the centre of
the hole pattern as possible and rapped solidly with a hammer. In either case, the tool-
maker should not attempt to obtain a deep centre impression for drilling. He should
not attempt to spot dowel holes.

Transfer Brass drift


punch Die shoe

Die block

Parallels
Shim
stock

Punch

Punch shoe Punch Transfer screw


(a) (b)

Fig. 2.55 Using transfer screws and transfer punches. The punches in (a) and (b)
are two different punches
The next step is to separate the assembly and drill the screw holes. Sometimes a
toolmaker enlarges the centre impressions with a centre punch, but unless he is
64 Tool Design

extremely careful, he cannot do an accurate job of drilling because the location of a


standard drill in large centre-punch hole is often an educated guess. A better method
is to locate directly from the centre impressions with a wiggler or locating microscope
and then start the hole with a spotting drill or centre drill, followed by a regular drill.
After the screw holes have been transferred, drilled, counterbored, etc., the punch
may be doweled in place. The punch is again positioned in the die opening, using
shims if required. Screws are inserted in the holes and snugged securely but not to
lock tightness. A small, soft drift and hammer are used to tap the punch as required
until positioned. The screws are then tightened all the way. This may disturb the punch
position slightly, and it is necessary to check after final positioning.
Final checking is accomplished in several ways, but commonly by using a light or
feeler gauges or by cutting a suitable material. Use of a light consists of shining a
small flashlight (often called a die light) into the junction of a closed punch and die
and observing the light line on the other side. This works very well when the die clear-
ance is extremely small, providing the die opening has a clearance opening in the die
shoe so that the toolmaker can see. For dies with clearance opening, the punch posi-
tion can be accurately checked by making trial cuts in a suitable material. Soft tissue
paper or cigarette paper works well for dies with small clearances; soft, heavier paper
is used for larger clearance. Soft paper tends to pull and tear on one side instead of
cutting clean when the punch and die relationship is incorrect. When the final position
is checked, the dowel holes may be drilled and reamed. Dowel holes were discussed
earlier in this chapter.
The previous discussion cites a typical example of the use of transfer methods for
locating holes. It is limited to situations like this and others when the rough alignment
of the holes in two workpieces is more important than their relation to dimensions.
Transfer methods work well for screw holes because the holes are somewhat over-
sized to allow movement of one component in relation to the other. This is not the
case when making dowel holes, which must be in near perfect alignment. Transfer
methods would not be accurate enough in this case. The transfer punch itself will not
consistantly transfer the centre point because the angle at which it is held when trans-
ferring is not always the same. The real problem, however, comes when the toolmaker
begins to align the transferred centre with the cutting tool to produce the hole. No
matter how carefully he works, there is almost certain to be an additional inaccuracy,
called transition error. Transition error is cumulative in direct proportion to the number
of operations required to pick up the location and inaccuracies of the machine spindle
and chuck. The wiggler and locating microscope are only as accurate as the chuck
and machine spindle. For example, the point of the wiggler is on the theoretical axis
of rotation and will locate the spotted point under the theoretical axis of rotation when
properly used. However, the chuck must hold the cutting tool on the same axis of
rotation, and if the chuck is not accurate, this is impossible. The chuck on most drill
presses is not this accurate.

Jig boring and jig grinding It was inevitable that the inefficiency, unreliability,
and dependence upon individual skill of the previously described toolmaker meth-
ods of locating holes and contours should eventually bring about the development of
mechanical aids engineered to solve locating problems. The ultimate in these devel-
opments has been the jig borer and its companion machine, the jig grinder.
Toolmaking Practises 65

In a general sense, the jig borer is similar to a milling machine. As with milling
machines, there are vertical-spindle machines and horizontal-spindle machines. The
vertical-spindle jig borer is more commonly used in toolroom work, and the discussion
will be limited to this type.
The basic elements and action of a jig borer are essentially the same as those of the
vertical milling machine. The cutting tool rotates on a spindle which is arranged for
feeding along a vertical axis. Positioning of the workpiece is accomplished in a hori-
zontal plane with longitudal and transverse feed movements.
Although there is a general similarity between the vertical-spindle milling machine and
the jig-boring machine, there are small differences that are extremely significant. The
most important is in the accuracy of positioning the workpiece. The modern jig borer
is capable of positioning to within 2.5 mm for absolute location or spacing of holes. The
spindle is also more rigid and precise to provide accurate concentric machining. The
jig borer is more massive in proportion to the capacity of the machine, to minimise
small dimensional errors due to deflection, vibration, and temperature change.
A jig grinder is a jig borer equipped with a high-speed grinding spindle, which is gener-
ally powered by a high-speed pneumatic turbine. The grinding spindle may be offset
from the axis of rotation and orbit in a planetary path to grind the desired hole. The
jig grinder is used in the same manner as the jig borer, with the added convenience
of being able to machine hardened steel. For example, holes that become distorted
during the hardening process can be brought to correct size and position with the
jig grinder. In addition to grinding straight holes, most jig grinders can automatically
grind tapered holes and combinations of radii, tangents, angles, and flats. The posi-
tioning system of a jig grinder is essentially the same as that of the jig borer, except
that provisions are made to protect the ways from abrasive dust. Somewhat greater
accuracy may be obtained in grinding, simply because this operation on hard material

Fig. 2.56 The Moore no. 2 jig grinder Fig. 2.57 The jig borer (The Moore
(The Moore Special Tool Special Tool Company)
Company)
66 Tool Design

is inherently better suited to dimensional control than the machining of soft metal. Jig-
boring and jig-grinding machines are shown in Figs 2.56 and 2.57.
Numerous attempts have been made to convert vertical-spindle milling machines to
jig borers. Superficially, it would appear that the addition of an accurate measuring
system would serve this purpose. However, several factors prevent the attainment by
a milling machine of accuracy comparable to that of a jig borer, even with an identical
measuring system. A milling machine with its gibbed, vertically adjustable knee lacks
rigidity. In addition, the general construction, spindle design, and alignment of posi-
tioning elements are not sufficiently good to transfer all the accuracy of the measuring
system to the work.
Conversion of a vertical-spindle milling machine to serve as a jig grinder involves even
greater difficulty. A high-speed grinding spindle must be mounted eccentrically on the
main spindle, and there is still the same need for a measuring system. Together, these
two additions provide an approximation of the basic rectangular positioning move-
ment and planetary grinding movement required for jig grinding. Due to the relatively
greater complexity and refinement necessary in true jig grinders, this type of conver-
sion is less satisfactory than that of a jig borer.
The measuring systems of jig borers and jig grinders utilise lead screws, mechani-
cal or electrical gaging, or optical measuring, precision lead screws being the most
common. Some jig borers make use of a compensating lead screw. Compensation is
accomplished by means of a cam which causes slight variations in the angular posi-
tion of the lead-screw nut. These variations offset any deviations in the lead screw
itself.
Mechanical measuring or gauging methods used on jig borers generally consist
of building up a set of gauge blocks or end measures to the required dimensions.
The built-up blocks or measures lie in a trough along the side of the table, and the
final fractional dimension is obtained by a micrometer adjustment or a dial indicator.
Recent development of this system utilised electrical gauging or measuring systems.
Optical measuring systems consist of an extremely accurate engine-divided scale for
each axis. The scales are generally mounted inside the machine to prevent damage.
They are read by an optical system that magnifies the graduations. Machines with this
type of measuring system use lead screws to impart motion to the table and spindle
and depend entirely on the scales for accuracy.
The significant features of the jig borer lie in its directness and rapidity in attaining
the highest order of locational accuracy. These attributes, together with its versatility
and efficient stock removal, explain its rapidly increasing acceptance in widely varied
phases of the mechanical industries, from toolroom to production lines, from experi-
mental laboratory to inspection department.
Since the jig borer was developed primarily to solve the toolmaker’s ever present
problems of precise location, its first acceptance and widest general use have been in
the toolroom. Used alone or in partnership with its companion machine, the jig grinder,
the jig borer has done much to promote interchangeability in toolmaking. It has also
increased the toolmaker’s capacity, relieving him of the tedious job of locating holes
by makeshift methods.
As its name implies, one of the basic uses of the jig borer is in jig making. Jigs may
range from the simple plate type, requiring only an accurate dimensional relationship
Toolmaking Practises 67

between bushing and locator, to complex box jigs incorporating numerous holes and
surfaces in several planes. Even in cases where the workpiece to be produced from a
jig does not require hole location to a “tenth”, the bushing holes can be jig-bored much
more quickly and to high accuracy than they can be produced by any other means to
even broad locational tolerances.
Press tools, ranging from simple dies not requiring the highest accuracy to compound
and progressive dies necessitating the greatest precision, are an obvious product for
the jig borer, as shown in Fig. 2.58. As a case in point, the significant accuracy contri-
bution of the jig borer to the stator-rotor lamination lies in the precision location of the
holes in the punch plate. Worthy of special mention are the pilot holes and the bushing
holes guiding the aligning posts of the stripper, as shown in Fig. 2.59.

Fig. 2.58 Typical jig-borer work is represented by these progressive dies (The Moore
Special Tool Company)

Fig. 2.59 Punch plate and stripper of a stator-rotor lamination die, illustrating the need for
locational accuracy of the jig borer (The Moore Special Tool Company)
68 Tool Design

Dies not requiring high accuracy can be jig-bored and left unground after harden-
ing. Distortion errors in hole location can be determined by picking up the location
of the holes after heat treatment by using the jig borer. This establishes a new set of
co-ordinates to which the unhardened, related members such as stripper and punch
plates can be jig-bored.
More complex or precise dies require grinding after hardening. The ideal solution,
both from the standpoint of efficiency and accuracy, is jig grinding. In this way, all
parts can be jig-bored to co-ordinate location, with all die openings left a few microns
undersize to permit re-establishment of location by jig grinding after hardening. Parts
not requiring heat treatment may be bored to size.
The advantage in jig boring parts which will subsequently be hardened and jig-ground
is purely a matter of efficiency. It is normally desirable to leave more stock for grinding
than is necessary for economic reasons. This condition can prevail only if the original
location is precise, because any tolerance at this point must be added to the allow-
ance for anticipated distortion. The fact that the jig borer can locate holes accurately
and remove stock so rapidly is all to the good, since there is no faster method within
acceptable limits of accuracy. In jig boring holes to be ground later, it is unnecessary
to size to a “tenth”, which is a further saving of time.
Assuming that a jig grinder is not available, the jig borer may be used to re-establish
co-ordinate location prior to grinding in a conventional internal grinder. This is accom-
plished in the following sequence:
1. After the die block is hardened and surface-ground, soft plugs (steel or brass)
are fitted to the holes to be ground.
2. The dies is set up in a jig borer and small holes bored in the plugs to the same
co-ordinates as used in the original jig boring; this re-establishes the desired
location, which could no longer be trusted after hardening.
3. One at a time, these holes in the plugs are indicated true with the axis of the
internal grinder, as shown in Fig. 2.60. The plug is then removed and the die
opening ground to size (see Fig. 2.61).

Fig. 2.60 Indicating the hole from the spindle of an internal grinder (The Moore Special
Tool Company)
Toolmaking Practises 69

While this is not as accurate as jig grinding


and far less efficient from every standpoint, it
is most satisfactory for substitute methods.

The coordinate locating system In use, the


jig grinder and jig borer may be considered
as consisting of a vertical spindle for machin-
ing and a horizontal table for work position-
ing. Positioning is accomplished through a
rectangular co-ordinate system. The co-or-
dinate locating system is a method of posi-
tioning work by rectangular movement of the
jig-borer or jig-grinder table with reference to
a common zero. The basis of the rectangu-
lar co-ordinate system lies in dimensioning Fig. 2.61 Grinding a hole in the con-
all points from crossed ordinates, or zero ventional manner after the
lines, external to the workpiece. Figure 2.62 plug is removed (The Moore
shows these as scales along the upper and Special Tool Company)
left-hand edges of a piece of drawing paper.
In Fig. 2.62 the workpiece is shown in drawn into the intersection of these scales, its
edges aligned with zero in both directions. In this presentation, the location of the
holes in relation to the scales and to the conventionally dimensioned drawing of the
same piece, shown in Fig. 2.63, is clear.
10.0000
11.0000

12.0000

13.0000

14.0000

15.0000
1.0000

2.0000

3.0000

4.0000

5.0000

6.0000

7.0000

8.0000

9.0000

0.0000
1.0000

2.0000

3.0000
4.0000

5.0000

6.0000
7.0000

8.0000

9.0000
10.0000

11.0000

12.0000

Fig. 2.62 Graphic representation of workpiece, crossed ordinates, and graduated scales.
In dimensioning, this relationship is based on imaginary scales and zero lines. In
machining, the scales are represented by the lead screws and the intersection of
the zero lines by the spindle axis (The Moore Special Tool Company)
70 Tool Design

29 51 55 35
1 1
64 64 64 64

3
1
32
53
64
43
64
51
64

Fig. 2.63 The cumulative nature of conventional dimensioning is more haphazard than co-
ordinate dimensioning (The Moore Special Tool Company)

Redrawing the same piece more nearly central on the paper (Fig. 2.64) in no way
alters the relationship of the holes to each other and to the edges, even though the
latter no longer are aligned with the zero ordinates. From this example, it may be seen
that the relationship of the zero ordinates to the workpiece is a matter of choice and
can be altered to suit requirements.
10.0000

11.0000

12.0000

13.0000

14.0000

15.0000
1.0000

2.0000

3.0000

4.0000

5.0000

6.0000

7.0000

8.0000

9.0000

0.0000
1.0000

2.0000

3.0000
4.0000

5.0000

6.0000
7.0000

8.0000

9.0000
10.0000

11.0000

12.0000

Fig. 2.64 Relocation of the workpiece within the co-ordinate system does not alter its
dimensional values (The Moore Special Tool Company)
Toolmaking Practises 71

The advantage of having the zero lines external to the workpiece becomes more
apparent in the case of a workpiece picked up from an existing hole or dowel, as
shown in Fig. 2.65. Should the pickup point be considered zero, it is apparent that
any locations to the left or above this point would be indicated by a confusing negative
value, incompatible with any measuring system.

+ 0.3125
+ 0.9375
+ 1.0468
– 1.2500
– 0.9531
– 0.6562

0.0000
– 0.3750

– 0.7500

– 1.5000

0.0000

+ 0.6250
+ 0.7500
+ 1.0000

Fig. 2.65 Inability to shift the relation of the pickup point of the work to the co-ordinate
system would result in confusing negative values, as illustrated in this example
(The Moore Special Tool Company)

Figure 2.66 represents a more complex piece than that shown in Fig. 2.63 with the
dimension rewritten in the form of a conventional jig-borer layout for operator conve-
nience. An alternative (though less effective) presentation is shown in Fig. 2.67.
While the ideal arrangement would call for all drawings to be presented to the operator
in the form of approved jig-borer layouts, this is seldom given sufficient consideration.
Several points favour this arrangement:
1. Once introduced, the rectangular co-ordinate system of dimensioning is often
as convenient in original design as the more conventional form.
2. Drawings which have been conventionally dimensioned usually require a cer-
tain amount of calculation in conversion to co-ordinates. This may necessitate
a knowledge of trigonometry and mathematical ability beyond that which might
be expected from an operator. This operation can therefore be performed more
efficiently and safely in an engineering office than in the shop. When more than
an occasional job of this sort is done, a calculating machine capable of extract-
ing square roots will justify its cost.
3. Calculations made by the operator not only use up his time but keep the
machine standing idle.
In order to apply the example of co-ordinate layout to the machine, the workpiece must
be set on the machine table and one edge zeroed in line with the spindle. (Methods
72 Tool Design

of zeroing are explained later in the chapter.) The zero bracket on the scale is then
set to the nearest full inch and the corresponding dial set to zero. The procedure is

4.875
5.437

7.187
3.500
3.750
4.000
4.250
4.312

5.500

6.000
6.250
6.562
6.812

7.500
2.312
2.625
2.875
3.062
3.375
3.625
3.875
4.125
4.187
4.562
4.812

5.437

Fig. 2.66 Preferred form of co-ordinate dimensioning (The Moore Special Tool Company) 7.1875
3.750
3.500

4.875

2.312 2.3125
4.312

2.625
4.000 2.875
6.250

3.0625
5.500

3.375
6.000
4.250

3.625
3.875
4.125 6.5625
7.500

4.187

4.562
5.4375

4.8125
6.8125

5.4275

Fig. 2.67 An alternate method of co-ordinate dimensioning


(The Moore Special Tool Company)
Toolmaking Practises 73

identical in relation to the other edge. On completion of this step, it may be neces-
sary to alter the figures on the layout to agree with the new settings of the scales and
dials. Reference to Fig. 2.55 will show that when new values are established during
pickup, correction of the locational values presents no problem. It requires only the
adjustment by whatever numerical difference exists between the pickup and the layout
reference. Figure 2.68 illustrates the relation of work to machine.

70
10

80
8
0
11.1875

90
0
7.3980

0 1 2 3 4 5 6 7 8 9 10 12 13 14 15

7
6
5
4
3
2
1
0

0 5 10

10 0 90
2.0000

0.2050
5 10
0

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14

0 5 10

0 90 80 70

Fig. 2.68 Relationship of reference scales and lead-screw dials with co-ordinate dimensions,
shown in two locations (The Moore Special Tool Company)
74 Tool Design

Should the starting point be an existing


– 1.750 R.
hole or dowel, it is not necessary to start
from an even inch; the scales and dials
may be set to the values indicated on the 8 Holes
drawing. In this case, it may be necessary 45° spaching
to convert for the inch values.

Polar co-ordinates This is the conven-


tional dimensioning system employed
to designate the location of holes in a
circle, equidistant from a common point,
Fig. 2.69 Polar co-ordinate dimensioning
as shown in Fig. 2.69. Such dimension-
of a circle of holes (The Moore
ing takes the form of an angular value
Special Tool Company)
between holes and their radius from a
common centre. In establishing the angle for each hole, it is necessary to divide the
number of holes into 360° in order to determine the spacing angle which must be
successively added to each hole. Should the number be unevenly divisible into 360°,
care must be used in the calculation, since even a small numerical error will result in
a sizeable cumulative error.
Once derived, polar co-ordinates lend themselves directly to use with the rotary table
for machining setup (see Fig. 2.70). By displacing the centre of rotation from the
spindle axis to the amount of radius given, each hole can be located successively by
angular setting of the rotary table.
The calculation of polar co-ordinates can
be simplified by the use of Woodworth
tables, which give the angle between
holes for any number of evenly spaced
holes up to 100. Woodworth tables were
developed by WJ Woodworth to enable
jig-borer operators to obtain the correct
figures with the least effort. The tables
are often listed in operators’ manuals
for jig borers. Probably the best-known
source of Woodworth tables is “Holes
Contours, and surfaces,” published
by the Moore Special Tool Company,
Bridgeport, Connecticut, USA. Fig. 2.70 The use of the rotary table (The
Moore Special Tool Company)
Unless a highly accurate rotary table
is used, errors of considerable magnitude may result from polar positioning. As a
result, it is often desirable, for the highest accuracy, to convert polar to rectangular
co-ordinates and use the measuring system of the jig-boring machine for spacing.
The strongest deterrent to this practise lies in the lengthy calculations necessary.
In order to simplify this, Woodworth tables also give factors for direct conversion to
rectangular co-ordinates.
Controversy over the relative merits of the polar and rectangular co-ordinate systems
in practise cannot be resolved without consideration of all the factors pertaining to any
specific job. Certain general points, however, can be cited in reference to work done
on a rotary table.
Toolmaking Practises 75

The following are the most important:


1. Picking up the centre of the rotary table in relation to the machine spindle pro-
vides a source of error.
2. Relating the workpiece to both the axis of the rotary table and the spindle intro-
duces another chance for error.
3. Errors in the angular-measuring element of the rotary table must be taken into
account.
4. Any shift in the location of the spindle axis, as would result from temperature
change, is a potential source of locational inaccuracy.
The errors from the above sources may be doubled in the work, as a result of rotation
of the table, in much the same way that eccentricity of a revolving member is shown
as a double value on an indicator. In addition, any errors of the rectangular position-
ing system must be included, since the distance from the centre must be measured
by it. The evidence adds up to the fact that no matter how accurate the rotary table,
higher accuracy can generally be attained by use of the rectangular co-ordinate sys-
tem of location. Contours, represented by the punch and die section of the stator-rotor
lamination die, are most effectively dimensioned by rectangular co-ordinates. In this
example (Fig. 2.71) both the punch and die are made to the same layout.

4.000
4.694
4.805
5.0761
5.218

9.687
9.9365
10.187
10.4055

Fig. 2.71 Co-ordinate layout of punch and die from the stator-rotor lamination die (The
Moore Special Tool Company)

2.6 JIG-BORING PRACTISE


Since complete automaticity is an unlikely prospect for machines performing non-
repetitive operations on dissimilar workpieces, the human element cannot be over-
looked in any discussion of jig boring. An operator’s working habits and his knowledge
of the principles involved play an important part in the results obtained from a jig
borer. True, there is little opportunity for him to coax more accuracy from the machine
than is built into it, but there is always the danger that the reverse may occur.
76 Tool Design

What follows is a general outline of sound operating practises. It is not possible,


naturally, to anticipate all combinations of requirements and conditions encountered
in specific jobs.
Setting up the workpiece on the machine table establishes its geometric relationship
to the measuring system and the spindle axis. Although a number of variations are
possible, these are the most common requirements:
1. Parallelism between the surfaces or axis of the workpiece and the machine
table and the alignment of one edge of axis of the workpiece with the direction
of table travel.
2. Angular inclination of the workpiece to table surface.
3. Workpiece rotatability mounted to permit angular spacing of holes, the axis of
rotation being parallel to the spindle axis of inclinable at any desired angle to
90°.
The proper positioning of the workpiece and its clamping, or restraint, against unde-
sired movement during machining constitute the first step in any jig-boring job. Typical
examples falling into case 1 will first be discussed.
In order to avoid machining into the table surface itself, it is usually necessary to sup-
port this type of workpiece on parallel pieces. Although there is no objection to the use
of accurately ground and paired parallels of conventional shape, the parallel setup
blocks (Fig. 2.72) frequently prove to be far more convenient, particularly since they
provide a large choice of heights and strategic placement of support, so that interfer-
ence with the position of holes is avoided, as shown in Fig. 2.73. These blocks are so
made that their matching dimensions are alike within extremely close limits, permit-
ting their use in combination and as matched sets. Convenient clearance and tapped
holes facilitate their attachment to each other and to the table and the direct mounting
of the workpiece when desirable.

Fig. 2.72 Parallel setup blocks are Fig. 2.73 Setup blocks can be positioned
available for the jig borer in to provide support for long
sets (The Moore Special Tool workpieces without interfer-
Company) ing with machining locations
(The Moore Special Tool
Company)
Toolmaking Practises 77

While it may seen obvious, the importance of clamping directly in line with, or over, the
support cannot be overemphasised. Any pressure exerted on an unsupported portion
of the workpiece will inevitably induce a twist or bow and the condition illustrated in
Fig. 2.74. Similar difficulty that is encountered is clamp work which is not flat.

Fig. 2.74 Spring induced by unsupported clamping will result in a nonparallel hole axes
(The Moore Special Tool Company)

The set of bolts, straps, and heel rests shown in Fig. 2.75 is designed to provide an
efficient means for clamping a wide variety of work. Their conventional use is shown in
Fig. 2.76, the brass heel rests straddling the table T slots and opposing the clamping
pressure without marring the table surface. Extension pieces are for use with higher
work, such as box jigs.

Fig. 2.75 Bolts, straps, and heel rests provide the most convenient means of clamping
(The Moore Special Tool Company)
78 Tool Design

Fig. 2.76 Effective clamping leaves the tool area unobstructed, holds work securely, and
does not mar the table (The Moore Special Tool Company)

The second requirement, i.e., alignment of one edge of the work with table travel, can
be met in several ways. A straightedge, parallel to the travel, is provided at the edge
of the table. Thee workpiece can be set directly against this or, if desired, spaced
away by parallel pieces such as gauge blocks or parallel setup blocks (Fig. 2.77).

Fig. 2.77 Setup blocks space work from the straightedge, parallel to table travel (The Moore
Special Tool Company)

The alternative method involves use of an indicator held in the spindle nose by means
of the indicator holder (Fig. 2.78). With the indicator point brought into registration
with the edge of the workpiece, partial rotation of the spindle by means of the Knurled
diameter of the holder quickly establishes the proper position of the indicator with
its movement normal to the edge of the piece. With the indicator in this relationship,
transversing the table will move the workpiece edge past the indicator point. In this
way the indicator will show the amount of misalignment; this can be eliminated by
gently tapping the proper end of the work until the indicator shows no movement dur-
ing the traverse. The clamping nuts should then be snugged down and the alignment
Toolmaking Practises 79

verified again to ensure against an unno-


ticed shift. The dimensional relationship of
work to the measuring system and spindle
axis in each situation will be taken up later
as a separate phase of the operations.
Angular inclination of the workpiece to the
table surface and spindle axis, case 2, is
most accurately adjusted with the micro-
sine plate (Fig. 2.79). Based on the familiar
sine-bar principle, this accessory provides
an inclinable surface of adequate propor-
tions for the attachment of even rather large
work. Any desired angle from 0 to 90° can Fig. 2.78 Indicating edge of work
be accurately established, selected by ref- parallel with table travel
erence to the simple formula on the plate. (The Moore Special Tool
A non-influencing clamp secures the setup Company)
against movement during machining.

Fig. 2.79 Use of microsine plate to establish angular position of work


(The Moore Special Tool Company)

Normally, the edges of the microsine plate and the work are both aligned with the
table travel. However, special circumstances may require the development of a com-
pound-angular relationship, in which case a rotary table is mounted on the microsine
plate. By attaching the work to the rotary-table top, any desired compound angle can
be attained by the combination of the two available angular movements (Fig. 2.80).
Undue wear of table surface and ways can be avoided by developing the habit of set-
ting up jobs on the ends and corners of the table as well as in the centre. In this way
wear is distributed, and often an emergency or rush job can be set up and machined
without disturbing the workpiece already in place.
Now the workpiece is geometrically oriented to the machine’s travel. The next step
required establishing its dimensional relationship to the measuring system and spin-
dle axis. Since the required reference may be to an edge, pin, boss, hole, scribed
80 Tool Design

line, slot, or counter, a variety of pickup


techniques must be employed.
In setting dials or positioning the table,
movement must always be made in the
direction indicated by the arrows on the
dials, in order to eliminate backlash. If it
is necessary to reapproach a setting, back
away from the setting by several thou-
sandths. Then make the setting again from
the correct direction.
The indicator set (Fig. 2.81) includes the
equipment necessary to meet most of
these conditions. The indicator itself is
supported by an adjustable, jointed arm, Fig. 2.80 The combination of rotary
from either the indicator holder or drill table and microsine plate
chuck held in the spindle (Fig. 2.82). The is employed in compound-
flexibility of this arrangement permits con- angle setups and for their
siderable latitude in positioning the indica- index spacing (The Moore
tor to suit the workpiece. Special Tool Company)

Fig. 2.81 The indicator set consists of an Fig. 2.82 The indicator holder permits a
indicator, a jointed indicator wide range of adjustment yet
holder, a line finder, and an edge remains firmly positioned as set
finder (The Moore Special Tool
Company)

The most common requirement–reference to an edge–can be accomplished most


easily and accurately with the edge finder (Fig. 2.83). This accessory is so con-
structed that its surface registering against the edge of the work is precisely central
to the edges of the 10 mm wide slot provided for pickup with the indicator. The edge
finder is held firmly against that work; the table is positioned so that the indicator
shows an equal reading when rotated to contact the opposite, inside edges of the
slot. At this point, the edge of the piece will be exactly in line with the spindle axis. A
known relationship is thus established to the measuring system. This can be checked
Toolmaking Practises 81

by moving the table 5 mm and indicating the edge of the piece itself (Fig. 2.84). In this
way, any failure of the edge finder to register properly, due to a burr or irregularity on
the work, will be detected.

Fig. 2.83 Picking up an edge with the Fig. 2.84 Indicating directly against the
edge finder and indicator (The workpiece after use of edge
Moore Special Tool Company) finder automatically checks the
pickup (The Moore Special
Tool Company)

Fig. 2.85 Picking up edge without an edge finder. The indicator is set against the edge of
the workpiece, raised, and rotated 180° to touch the gauge block held against
the edge (The Moore Special Tool Company)

A somewhat less convenient and less accurate method picks up an edge without an
edge finder. Reference to Fig. 2.85 will show that the necessary vertical movement
exaggerates any edge deviation.
82 Tool Design

Holes, pins, bosses, and radiused contours are generally capable of single and direct
pickup by means of an indicator. A scribed line or point requires a different instrument.
Usually, the relative inaccuracy of such a reference makes it unnecessary to use a
more refined device than the line finder or wiggler shown in Fig. 2.86. Held in the rotat-
ing spindle, its point can be made to run true by a touch of the finger. It is then brought
close to the line or mark on the work, which is positioned directly under the point as
observed through a glass. This method may also be used for locating an edge which
may not, in itself, be sufficiently accurate to justify greater precision of location.

Fig. 2.86 Picking up the end of the die block with a line finder (The Moore
Special Tool Company)

Occasionally, the reference point on the work will present conditions not readily suited
to use the indicator and requiring greater accuracy than is attainable with the line
finder. Very small or partial holes, irregular contours, slots, and punch marks fall
into this category. A satisfactory solution is available in the locating microscope (Fig.
2.19). This instrument combines several unusual features essential to its use for this
purpose:
1. It includes a roof prism which reverts the image to its natural position, so that
positioning is a normal movement to the operator.
2. Its field of vision is large enough, to encompass an adequate area of the
work.
3. The 40X magnification is sufficiently great to permit a “tenth” to be seen.
4. The reticle reference consists of a number of concentric circles and two pairs
of crossed centre lines, suitable for a wide variety of requirements.
5. It has an adjustable optical axis which can easily be brought into coincidence
with that of the spindle.

Machining It is in the machining operation itself that the widest choice of methods
is available. It is seldom possible to establish a clear superiority of one over the other.
The choice of cutting-tool systems may be compared to the choice between roads,
Toolmaking Practises 83

all of which converge at a common des-


tination. This goal represents the highest
order of geometric, locational, and dimen-
sional accuracy, attainable only through
single-point boring. Should something
less accurate suffice, it merely means
stopping short on one of the chosen
roads. An attempt will be made to show
this in practise by the following discus-
sion of tools and techniques.
The usual practise in starting a hole is to
employ a centre drill. For holes to be fin-
ished by end reaming or those too small
to bore, a special spotting tool (Fig. 2.87) Fig. 2.87 Special spotting tools provide
should be employed to provide higher ini- high initial locational accu-
tial accuracy. The three sizes of special racy, important where hole is
jig-borer drill chucks (Fig. 2.88) are suit- not to be finished by single-
able for holding these tools, as well as the point boring (The Moore
various drills required in each operation. Special Tool Company)

Fig. 2.88 Precision drill chucks for jig boring (The Moore Special Tool Company)

Holes smaller than about 1.6 mm can be finished with satisfactory accuracy by drilling
directly from the spotted location. Larger holes can be rapidly opened up to within a
few thousandths of finish size, using the minimum number of drills in the process.
To extend the range of standard drills below 12 mm which should be available with
each machine, and because straight-shank drills over 12 mm are too long to use in
a jig borer, a special series of short drills in sizes from 12 mm to 40 mm has been
established, as shown in Fig. 2.89. Possible damage to the finely finished 12 mm
end reamer collet is avoided by making the shanks of these drills larger in size so as
not to enter the collet, thereby ensuring the use of the special chuck provided for the
purpose.
84 Tool Design

Fig. 2.89 Special jig-borer drills and spotting tools (The Moore Special Tool Company)

Roughing can be continued beyond the range of available drill sized by a variety of
boring tools. The Hole Hog, designed of this purpose, is particularly efficient in its rate
of stock removal in larger holes. A series of similar sweeping tools (Fig. 2.90) is equally
efficient in rapidly enlarging holes or in sweeping or facing operations on surfaces and
bosses. Pilots of sweeping tools are made to fractional sizes and, being hardened and
ground, will not damage a previously finished hole when used to sweep a boss.

Fig. 2.90 Sweeping tools for facing and rapid hole enlargement
(The Moore Special Tool Company)

Spotters, drills, and end reamers are shown in Fig. 2.91; when held in a collet, they
provide the fastest system of locating, drilling, and reaming holes on a jig borer. The
spotters establish locational accuracy, and the special undersize drills leave from
130 mm to 400 mm for final sizing by the end reamer.
Toolmaking Practises 85

Fig. 2.91 Special spotters, drills and end reamers used in jig-boring operations
(The Moore Special Tool Company)

Before passing to the subject of intermediate and/or finishing cuts, it might be well to
point out that these generally account for most of the time per hole. Therefore, in the
interest of efficiency, it is advisable not to start them too soon, i.e., rough to within as
close to finish size as practical.
Although not matching the accuracy of the slower single-point boring, two types of
end reamers provide a favourable compromise, i.e., a considerable saving of time at
an often permissible minor sacrifice in accuracy. The end reamer, held rigidly and run-
ning true with the spindle, acts somewhat like the combination of a boring tool and a
reamer, locating and sizing at the same time. This hybrid tool does not achieve perfec-
tion in either respect, for a slight runout tends to make it cut slightly oversize. Gently
clipping the corners of the teeth with a stone improves this condition but somewhat
reduces the end reamer’s ability to establish location. With reasonable care, however,
an operator should be able to hold size and locational errors to within ±130 mm.
End reamers are available in sets of various sizes as shown in Fig. 2.92. Some spe-
cial set of drill reamers are also available for sizing holes which are needed to be
ground.
End reamers are held in collets (Fig. 2.93), the locating bore of which is ground and
lapped to a plug-gauge fit to the reamer shank. For the highest accuracy, it is some-
times desirable to peen the reamer shank slightly to make it run true in the collet,
indicating from the teeth.
Ordinary rose or fluted reamers differ from end reamers in the relative flexibility of
shank, which permits them to align themselves with the hole. They are often a sat-
isfactory means for sizing a hole (Fig. 2.94). When carefully handled and removing
between 25 mm and 80 mm of stock, these tools will produce somewhat better holes-
size accuracy than the end reamer. Either type of reamer has the advantage of avoid-
ing the cut-and-try adjustment necessary with a single-point boring tool, but they can-
not produce the infinite range of sized can.
86 Tool Design

Fig. 2.92 Straight-shank end reamers (The Moore Special Tool Company)

Fig. 2.93 Method of holding end reamers Fig. 2.94 Rose or fluted reamer reaming a
in collets (The Moore Special previously bored hole to size (The
Tool Company) Moore Special Tool Company)

The importance of single-point boring as the most accurate method of attaining loca-
tional accuracy in jig boring amply justifies the wide range of boring tools available.
Size control in single-point boring depends on a means of adjusting the position of
the tool point outward from the axis of the spindle rotation. Two types of boring chucks
provides this adjustment through different means. The swivel-block type (Fig. 2.95)
offers the widest range of adjustment for its diameter, a decided advantage in working
in close quarters, as in deep box jigs or close to die posts.
A disadvantage frequently noted in this type of chuck lies in the fact that adjustment
of the tool moves it in an arc; thus the graduation of the adjusting screw does not bear
and exact relation to the tool movement. As a matter of fact, this is not so serious a
problem as might be anticipated since the graduations (quite accurate over a short
range, provided tools of proper length are used) would hardly be depended upon over
a long range of travel.
Toolmaking Practises 87

Fig. 2.95 A boring chuck of the swivel-block type (The Moore Special Tool Company)

From a purely functional standpoint, the extreme rigidity resulting from a large clamping
surface, together with the inherently clean design, makes it very reliable in operation.
The dovetail offset type of boring chuck (Fig. 2.96) is somewhat shorter than the
swivel block. As an added advantage, the tool is moved directly outward in a straight
line by the adjusting screw. This permits the rise of odd-length tools without altering
the value of the graduation, although the range does not equal that of the swivel type.
Due to this same straightline movement, this chuck is convenient for squaring shoul-
der and facing bosses.

Fig. 2.96 Boring chuck of the dovetail offset type (The Moore Special Tool Company)

Because of the necessary cut-and-try method of adjusting any boring tool, there is
definite advantage in being able to leave the adjustment set to cut to a particular size
for repetitive work. To this end, a series of boring tools (Fig. 2.97) has been developed.
These boring bars are of the solid type and are adjustable over a relatively short range
by means of graduated screws. The series covers a range of hole sizes from 12 mm to
90 mm. They are particularly useful in deep-hole boring (Fig. 2.98).
88 Tool Design

Fig. 2.97 Solid-type, thousandth-setting boring bars (The Moore Special Tool Company)

Fig. 2.98 Jig-boring deep hole with solid boring bar (The Moore Special Tool Company)

Carbide tools are very efficient for some types of jig-boring operations, particularly in
boring aluminium and cast iron or for production work. Therefore, in addition to the
set of high-speed tool bits shown in Fig. 2.99, duplicate sized are available tipped
with carbide.
Variation in the machinability of commonly encountered materials, together with the
difference in spring or deflection of various types of boring tools, make experience
the only practical way of mastering the technique of boring to size. Implied in any
discussion of finishing and sizing, however, is the attendent problem of measurement.
The ability to measure accurately is the most important factor in the ability to size
accurately.
Toolmaking Practises 89

Fig. 2.99 Boring bits for chucks, boring bars, and sweep tools
(The Moore Special Tool Company)

Measurement Relatively accurate measurement of hole diameter can be made with


the familiar inside caliper. When it is set to touch the sides of the hole, the dimensions
can be read on a micrometer. Slightly more accurate results may be had by substitut-
ing a telescoping gauge for the caliper.
More accurate than either of these is the plug gauge, but it can only measure a hole
whose size is exactly that of the gauge. Unless an infinite range of gauge sizes is
available, it cannot measure a hole before it is finished, nor can it determine how
much stock remains to be removed. To overcome this limitation, a hole-measuring
instrument, the flat-leaf taper gauge (Fig. 2.100) has been developed. This set of
gauges has graduations at the end of each overlapping those of the next.

Fig. 2.100 Leaf taper gauges (The Moore Special Tool Company)

This type of gauge contacts the hole at the top and, being short, conveniently clears
cutting tools. Obviously, it cannot determine errors in the geometry of the hole, i.e.,
bell mouth or taper, nor can such gauges be read to a “tenth”. They are invaluable as
a guide to approaching finished size prior to plug fit.
Also capable of more flexibility in use than a plug, and of equal accuracy, is the
indicating-type gauge (Fig. 2.101). Set against a micrometer or master ring, it readily
shows deviation from true cylindricity as well as the diameter of the hole.
90 Tool Design

Fig. 2.101 Indicating-type hole gauge are not only more versatile than plug gauges because
of their greater size range but also reveal the geometry of a hole (The Moore
Special Tool Company)

General operating practises While it is impractical to follow any prescribed set of


rules in the operation of a jig borer, due to the wide variety of work sure to be encoun-
tered, the following general sequence should be followed whenever possible:
1. Set up the workpiece carefully to ensure proper relation to machine travel.
2. Establish the dimensional relationship between the reference point on the work
and the spindle axis. Relate this to the measuring system by setting the scales
and dials as previously described.
3. Spot the position of all holes lightly with a centre drill.
4. Repeat step 3, respotting to the depth necessary to provide and adequate
guide for the subsequent drilling operation. While step 3 may appear unneces-
sary, the respotting is good insurance against most common operator error of
misreading the scales.
5. Rough all holes to nearly finish size before finishing any of them. Drilling or
sweeping are the most efficient means of accomplishing this operation. Not
only is the stock removed more rapidly than by boring, but less side thrust is
exerted on the work, an insurance against shifting. When drilling large holes,
the size should be increased in steps to avoid pushing the inefficient centre
web of large drills through solid stock.
6. Go back and check the original setting of the work, step 2, to make sure it
has not shifted during roughing. On particularly accurate work its temperature
should be allowed to return to that of the room before making this check.
7. Finish-bore all holes to size. The final cut, unless an excessive amount of stock
has been left from roughing, will not materially affect work temperature.
8. Check the location of each hole after re-establishing the original reference.
Indicating each hole from the spindle is an extremely sensitive and accurate
method of inspection, since the reading shows twice the actual displacement
or error of location.

2.7 INSTALLATION OF DRILL BUSHINGS


Drill bushings and their relationship to drill-jig design are discussed in detail in
Chapter 7. The information given is from the viewpoint of the tool designer, who
Toolmaking Practises 91

generally is not responsible for the actual installation of drill bushings. The following
information deals with drill bushings from the viewpoint of the toolmaker, who is gen-
erally concerned with bushing installation. The tool designer should be aware of the
bushing-installation procedure in order to understand the problems of the toolmaker.

General bushing installation data Generally speaking, the toolmaker should pro-
vide sufficient clearance between the end of the bushing and the workpiece to permit
the removal of chips. In other words, the chips should not be allowed to pass through
the bushing. The exception to this rule occurs in drilling operations requiring maxi-
mum precision, where the bushings should be in direct contact with the workpiece.
However, suitable chip clearance should be provided in most applications because
the abrasive action of metal particles will accelerate bushing wear.
The recommended chip clearance for metals (cast iron, for example) which produce
small chips should be equal to one-half the bushing ID. For other metals (such as
cold-rolled steel) which produce long stringy chips, the clearance should be at least
equal to the bushing ID but should not exceed 1.5 times the ID. Excessive chip clear-
ance should be avoided because cutting tools are slightly larger in diameter at the
cutting end and excessive clearance reduces the guiding effect of the bushing and
results in less accurate drilling (see Fig. 2.102).

Cutting tool
Excessive with normal
chip back taper
clearance Guiding
effect
Jig of bushing
ID reduced

Equal to: Workpiece


Workpiece 1 ID (small chips)
2
1 to 1 1 ID (long, stringy chips)
2
(a) (b) (c)

Fig. 2.102 Recommended clearance between workpiece and bushing: (a) No clearance
(maximum-precision drilling only) (b) Normal chip clearance (c) Excessive
chip clearance (American Drill Bushing Company)

When performing multiple operations, such as drilling following by reaming, slip


renewable bushing of different lengths may be used to obtain the combined advan-
tages of adequate chip removal and precise accuracy. The slip renewable bushing
should be short enough to provide proper clip clearance during the drilling operation,
while the reamer bushing may be long enough to contact or closely approach the
workpiece, thus providing maximum guiding effect during the reaming operation (see
Fig. 2.103).
Burr clearance should be provided between the bushing and workpiece when drilling
wiry metals, such as copper. Metals of this type tend to produce minor burrs around
92 Tool Design

Slip renewable
used for drilling

Jig
Jig
Burr
clearance

Phantom view Workpiece


shows slip
Workpiece renewable Secondary burr
reamer bushing (minor)

Fig. 2.103 Bushings of different lengths Fig. 2.104 Burr clearance should be
may be used for maximum provided when drilling wiry
guiding of multiple opera- metals (American Drill Bush-
tions (American Drill Bush- ing Company)
ing Company)

the top of the drilled holes. This is likely to lift the jig from the workpiece and cause
difficulty in the removal of workpieces from side-loaded jigs. The recommended burr
clearance is equal to one-half the bushing ID (see Fig. 2.104).
Extra-thin-wall bushings will prove helpful for many applications requiring close
centre-to-centre bushing placement, as shown in Fig. 2.105. However, for especially
difficult close hole patterns it may be necessary to grind flats on the bushing ODs and
/or heads to achieve minimum spacing. When this technique is used, the bushing flats
and mounting holes must be accurately machined.

(a) (b)

Fig. 2.105 Methods of obtaining close hole patterns: (a) Extra-thin-wall bushings with ODs
less than those of ANSI bushings of same ID (b) Standard headless or headed
bushings with ground flats on ODs (American Drill Bushing Company)

Figure 2.106 shows various methods of adapting bushings to applications involving


irregular work surfaces. Here the ends of the bushings should be formed to the con-
tour of the workpiece. In many applications of this nature, the drill point does not enter
perpendicular to the work surface and has a tendency to skid or wander. For this rea-
son, the distance between the bushing and the workpiece must be held to a minimum
so that the full guiding effect of the bushing can be obtained. The side load exerted by
the drill in an application of this type is usually concentrated at a point near the drill-
exit end of the bushing and causes accelerated bushing wear. Except in production
runs, the use of fixed renewable bushings should be considered to simplify replace-
ment of worn bushings and to facilitate proper orientation of the bushing with respect
to the contoured work surface. When press-fit bushings are used, bushing contours
should be applied after bushings are installed in the jig plate to assure proper place-
ment with respect to the workpiece.
Toolmaking Practises 93

Fixed Fixed
renewable Headless press-fit renewable
Jig bushing bushing, used for Lock screw bushing Liner
short production run

Jig Sloped
Shaft workpiece Workpiece

(a) (b) (c)

Fig. 2.106 Modification of drill bushings for irregular work surfaces: (a) End of bushing
formed to workpiece contour (b) End of bushing chamfered to conform to slope
of workpiece (c) Fixed renewable bushing used for high-production application
(American Drill Bushing Company)

Conventional installation of press-fit and linear bushings One of the important


factors in the installation of a press-fit bushing is fitting the bushing to its mounting
hole with the correct interference fit. Too much interference will cause closure of the
bushing ID or distortion of the drill jig plate, or both; too little interference will result in
a loosely installed bushing. To arrive at a correct fit, such variables as the material of
the jig plates, the thickness of the jig plate (or length of the mounting hole), and the
thickness of the bushing wall must be carefully considered. Because of these and
many other variables, no definite rules can substitute for the skill and judgment of the
experienced toolmaker. However, the following suggestions, given by the American
Drill Bushing Company, will prove helpful to beginners. These recommendations are
based on the use of cast-iron or unhardened-steel jig plates.

Avoid Excessive Interference Fits In any press-fit installation, metal is displaced.


Usually a combination of bushing closure and jig-plate distortion results; therefore the
best practise is to use the minimum interference necessary to retain the bushing in
the jig plate. In most installations, a diametral interference of 12 to 20 mm is adequate
for installation of headless press-fit bushings or liners.
For example, if a bushing of 12 mm ID by 20 mm OD by 20 mm length is pressed in
a 20 mm thick jig plate with an interference fit of 15 mm, the ID of the bushing will be
reduced by approximately 5 mm. This fit would be ideal since the bore of the bushing
is manufactured slightly oversize with a plus tolerance of 2.5 to 12 mm larger than the
nominal drill size. The slight decrease in the size of the bushing bore at installation
reduces the bushing ID almost exactly to the desired nominal ID and holds distortion
of the jig plate to a negligible amount.
A lesser interference fit of 7.5 to 12 mm can be used for installing head press-fit bush-
ings and liners. The head prevents dislodgement of the bushing caused by accidentally
striking the bushing with the drill during a production operation. The headed bushings
are particularly recommended for installation in comparatively thin jig plates, where an
excessive interference fit would be required to retain headless bushings. The headed
bushing also provides a greater drill bearing length in this type of installation.
An excessive interference fit may reduce the diameter of the bushing bore to less
than a practical work clearance. This can cause tool seizure or, in the case of linear
94 Tool Design

bushings, prevent insertion of a renewable bushing. The usual remedy for this con-
dition is to relap the bushing in the jig plate. This operation is both costly and time-
consuming and often results in a bell-mouthed bushing with a less effective bearing
length and a greatly shortened wear life.

Lubricate hole and bushing Before installing a press-fit bushing, always lubricate
the ID of the mounting hole and the OD of the bushing. White lead is an excellent lubri-
cant for this purpose. Bushings installed without lubricant may pick up metal and score
the mounting hole during installation. Lubricated bushings are more easily removed
for replacement with less chance of damaging the jig plate.

Install bushings with an arbour press whenever possible The preferable method
of installing press-fit bushings is with an arbour press, but when the jig plate is too
large for an arbour press, alternate methods must be considered.
One such method is used when the ID of the bushing is large enough to allow inser-
tion of a bolt. The bushing is first started into its mounting hole; then drilled plates
large enough to span the mounting hole, are placed on top of the bushing and on the
back side of the jig plate. A bolt is inserted through the plates and the bushing and is
fastened by a nut. As the nut is tightened, the bushing is safely and effectively pressed
into place.
When the bushing is too small to permit this procedure, a hammer and a soft-metal
punch of aluminium or brass may be used to carefully drive the bushing into place.
Never use direct hammer blows on the bushing; the bushings are heat-treated to a
high degree of hardness and may be fractured by such pounding.
A new method for installing bushings and liners with press-fit security involves the use
of American Loctite, a specially formulated bonding agent. In this method, the bush-
ings or liners are coated with the special bonding agent and are installed in mounting
holes that have been prepared to provide a close slip fit, instead of an interference fit,
thus preventing distortion of the bushings or jig plates. The security of installation has
been found equal or superior to normal press-fit installations.
Whatever method of installation is used, remember, that most drill bushings feature a
concentric-ground lead that will slip into the mounting hole and establish initial align-
ment. Always maintain this alignment until the bushing is well started into its mounting
hole.

Typical installations of renewable bushings Renewable bushings are usually


installed in liners and are secured in place by lock screws or clamps. In most applica-
tions, the liners (head or headless type) are mounted flush with the jig plate; however,
projected mounting is sometimes used for head liners when the jig plate is too thick to
accept a suitable counterbore or when the intended application does not warrant the
machining costs involved.
Lock screws are suitable for use with flush-mounted liners, usually in light-duty appli-
cations. For heavier duty applications, clamps provide a better means of locking the
bushing against the effects of vibration and torque produced by drill rotation. Clamps
provide a larger bearing surface against the jig plate and are secured by standard
socket-head cap screws.
Recommended methods for locking renewable bushings in flush or projected liners
are shown in Table 2.1.
Table 2.1 Typical installation of renewable bushings
Projected mounted
Flush-mounted (head or headless linears) (head liners only)
Fixed renewable bushings

Lock screw: Round clamp: Round end clamp:


recommended provides better security for locking standard fixed
for light drilling applications; than lock screw in heavier- renewable bushings in
small head diameter permits -duty applications; diameter projected mounted liners
close bushing placement same as lock-screw head for
close bushing placement
Round end clamp: Flat clamp: Flat clamp:
for locking standard fixed for type FX bushings only; for type FX bushings only;
renewable bushings; provides provides maximum security provides maximum security
large bearing surface against against vibration and torque against vibration and torque
jig plate

Slip renewable bushings

Lock screw: Round end clamp: Round end clamp:


recommended for light for heavy-duty applications for any applications using
drilling operations; small provides large bearing slip renewable bushings
head diameter per-mits close surface against jig plate installed in projected
bushing placement mounted head liners
Toolmaking Practises 95
96 Tool Design

2.8 PUNCH AND DIE MANUFACTURE


The manufacture of punches and dies shaped in common geometrical patterns, such
as rounds, squares, or rectangles, present no problem to the toolmaker. Standard
punches and dies are available commercially in the smaller sized and thus eliminate
the necessity of making them in the majority of cases. Installation is simply a matter
of jig-boring a hole pattern in the punch plate and a matching pattern in the die holder
and inserting the purchased punches and dies.
It is another matter when it comes to manufacturing punches and dies with shapes
that are irregular and do not have true geometric lines. Machine tools are unable to
impart the necessary shape on the punch and die, and the manual skill and ingenuity
of the toolmaker become important. However, modern machine methods have made
the toolmaker’s task easier and faster. For example, there was a time when the only
practical way of making irregular shapes and holes in thick sections of tool steel was
to drill a row of holes adjacent to a layout line as close together as possible without
overlapping. After completing the hole pattern all the way around the contour, the slug
was cut out by cutting the narrow strip of metal between each hole by a hammer and
cold chisel. The jagged edge was then smoothed with a file. The final finishing to the
layout line was also done with the file. This “blacksmith method” has been replaced
with modern band-sawing techniques which save the toolmaker countless hours of
manual work. However, the toolmaker is still working to the layout line, and the final
result still is determined by his manual skill.
This section discusses a few of the techniques and machines used in the manufacture
of punches and dies. The study of this discussion will not make a toolmaker out of the
beginning tool designer, but it will certainly help him appreciate the toolmaker’s prob-
lems. The techniques may be used singly or combined, depending upon the type of
equipment available. They are by no means complete and should serve to give ideas
which may be expanded by the individual.

Contour sawing of punches and dies Contour sawing on the metal-cutting band
saw is the most common method of roughing out irregular shapes in punches and
dies. With good blades and sawing technique, the toolmaker can saw within 0.4 to
0.8 mm of the finished contour. Die openings are sawed first by drilling a small hole
inside the scribed die opening and inserting the cut band-saw blade through the hole.
The blade is rewelded together and inserted into the machine, and the slug is sawed
from the die block. The blade must again be cut to remove it from the die opening. The
table of most band-sawing machines can be tilted in two planes to allow the sawing of
die clearance directly into the die blocks.
The contour-sawing process also enables the toolmaker to produced both the punch
and die from the same piece of tool steel. By machining out the slug with the contour
saw and cutting at an angle determined by the die thickness, the slug can be used as
a punch. In addition to eliminating the time ordinarily required to rough out the punch
and die there is a saving of the material that would have been used for the punch. The
procedure is as follows.
The die opening is scribed on the die block by conventional methods. An inner line
is then scribed on which the saw is to travel. The distance of this line inside the die
layout is determined by the angle of the cut, as shown in Fig. 2.107. It is started inside
Toolmaking Practises 97

Fig. 2.107 Internal contouring band sawing used for machining a blanking punch and
die [From ASTME, “Tool Engineers Handbook,” 2d ed., F W Wilson (ed.)
McGraw-Hill, New York, 1959, by permission]

the die-layout line, intersects the die-layout line at the centre of the die block, and
emerges through the die block on the opposite side, outside the die-layout line. The
drilling angle and diameter are determined by the thickness of the block and from
Table 2.2.

Table 2.2 Dimensional data for diemaking technique shown in Fig. 2.107
Die Angle of Angle Distance Distance from Diam Width of Amount of
thickness, saw-start- for saw from die die layout if drill, starting straight sides
mm ing hole, cut, deg layout line to line to centre mm saw, mm on punch
deg centre of saw of starting and die, mm
kerf, mm hole, mm

12.5 21 18 2 2.4 3.2 2.4 4.8


19 18 15 2.4 3.2 3.2 2.4 7.1
25.4 14 11 2.4 3.2 3.6 3.2 9.5
32 12 9 2.4 3.2 3.6 3.2 11.9
38 11 8 2.8 3.6 3.6 3.2 14.3
51 10 7 3.2 4.8 5.2 4.8 20.7
76 9 6 4 6.4 6.8 6.4 28.6
102 8 6 5.6 7.1 6.8 6.4 41.3
127 7 6 6.4 7.9 6.8 6.4 57.2
153 6 5 6.4 7.9 6.8 6.4 63.6
SOURCE: ASTME, Tool Engineers Handbook, F W Wilson (ed.), McGraw-Hill, New York, 1959, by permission
98 Tool Design

The narrow saw blade is then inserted through the die-starting hole, and the ends of
the saw are welded together. The machine table is tilted so that the table angle is a
few degrees less than the angle in which the hole has been drilled. The slug is com-
pletely sawed out at this angle. The path of the saw on the surface is entirely inside
the layout line. Thus at the proper cutting angle, the bottom of the slug that is to be
removed will have material outside the die-layout line.
When the punch is removed, the excess material on both the punch and die is filed
off to make the land on the die and the straight sides on the punch. When the correct
sawing procedure is followed, the starting hole will completely vanish.
The filling operation is best done on a filling machine. When the contour is open, a
continuous-band file is the most accurate and fastest method (see Fig. 2.24). Some
vertical band saws have provisions for inserting band files; however, a vertical band
filling machine is usually available in most toolrooms. Sharp corners and intricate
shapes may have to be finished on the die-filling machine shown in Fig. 2.26.
Matching contours of punches and dies. While contour sawing is an efficient method of
roughing out the general shape of punches and die openings, there is still a consider-
able amount of hand fitting that must be done before the die is complete. The method
of fitting to be described has been in use for a long time, and first requires making a
template that conforms to the shape of the blank which the die is to produce. If the die
clearance is to be applied to the die opening, it should be added to the template. This
template is then used as a gauge when finishing the opening in the die.
Cold-rolled steel from 1.6 to 3.2 mm thick, depending upon the size of the blank, is
used to make the template. The outline of the template is laid out carefully, as the
finished template will serve as the master pattern when fitting the punch and die.
The template is rough-sawed on the contour band saw and finished to the layout line
with a die-filling machine or hand-filling methods, depending upon the contour of the
template. Some toolmakers prefer to make the template from tool steel, which can
be hardened after band sawing. The template is then surface ground on both sides
and accurately laid out again. It is then finished by a rotary profile grinder, as shown
in Fig. 2.31.
A handle is generally fitted to a template either by drilling and tapping a hole in the
template and screwing a threaded handle into the hole or by attaching a wire with soft
solder.
Although the master template is used primarily for final finishing of the die opening,
it may be used for laying out the die. This is accomplished by clamping it in position
on the face of the blue die block and scribing around it with a sharp scriber. When
the opening must be located in relation to other holes, as with a progressive die, the
template position can easily be checked before scribing by using precision measuring
instruments.
After the die opening has been band-sawed to approximately the required size and
shape, the exact size and shape ordinarily is obtained by filling, by hand or on a filling
machine. Unless angular clearance was obtained during the sawing operation, the
first step should be to tilt the table of the filling machine to the angular-clearance angle
and rough out the angular clearance until a land of the proper width below the cut-
ting edge is obtained. The narrow land is much easier to precision-file because less
Toolmaking Practises 99

material must be removed. The land area is now filed until it fits the master template
nicely. The template is frequently tried in the hole during filling, and the bearing points
to be removed are marked with a pencil. Some toolmakers prefer to coat the edge
of the template lightly with prussian blue to mark the bearing points, which are filed
until the blue disappears. This operation is repeated until the hole fits the template
perfectly and will just allow it to pass through.
Final filling to shape and size requires a technique that is developed largely through
experience. To file a narrow surface straight is difficult, and many toolmakers use file
guides to prevent overfilling beyond the contour lines. Files guides are usually hard-
ened disks and parallels which are clamped to the die block or punch face, as shown
in Fig. 2.108. The disks and parallels are made from hardened tool steel (RC63 to
65) and are from 3.2 to 4.8 mm in thickness. The disk diameter ranges from 32 mm
to the largest size used by the toolmaker in increments of 0.80 or 1.6 mm. Parallels
are made in various lengths. The diameter of the disk should be marked directly on
the disk.
In use, the disks and parallels are clamped in position on the die block to form the
geometric contour of the die opening. Disks are used primarily for finishing convex
surfaces, while the parallels are used to finish straight lines (see Fig. 2.108). Concave
surfaces are more easily formed with single-point boring tools. Straight tangent line
between concave surfaces are finished by clamping parallels tangent to the concave
surfaces. The toolmaker files down to the file guide, being careful to avoid continu-
ously striking the hard edges.

Fig. 2.108 Use of hardened file guides to prevent overfiling

The amount of filling may be reduced, and in some cases completely eliminated, by
the use of a rotary cross-slide milling head attached to a vertical milling machine (see
Fig. 2.109). This attachment permits accurate milling of combinations of angles and
radii and angles tangent to radii and blending radii into radii with a single setup. The
toolmaker need not reset and re-align the workpiece for each radius to be machined,
as with the conventional rotary table. The main feature is the cross-slide motion,
which can be set off-centre and rotated in a continuous 360° planetary motion.
100 Tool Design

Fig. 2.109 The rotary cross-slide milling head (Volstro Manufacturing Co.)

The rotary cross-slide milling-head attachment can machine almost any geometric
figure that can be constructed with drafting instruments. To grasp the principle on
which it operates, consider the milling machine and rotary head as one large drafting
machine. The milling-machine longitudinal table would correspond to the horizontal
lines produced by the horizontal arm of a drafting machine. Similarly, the vertical lines
would be produced by the milling-machine cross movement. Offsetting the rotary
head spindle with the head cross slide and rotating the head would be like using a
compass to duplicate in metal various arcs, radii, and circles. The head cross slide,
in conjunction with angular settings of the rotary head, can also produce angular cuts
which correspond to the protractor on a drafting machine. Thus, the toolmaker, within
the limits of the range of available movements, is duplicating on the die block the
same movements made by the draftsman in producing a drawing with conventional
drafting instruments.
When the die is completed, the toolmaker is faced with the problem of machining the
punch contour to match that of the die. Round punches or punches whose contours
are composed of straight lines and true radii can be machined directly to size with
rotary milling tables or rotary cross-slide milling heads. However, irregular contours
may have to machined directly from a layout line and carefully handworked to fit the
die opening.
The customary practise is to harden the die and then make the punch to fit the die.
The end of the punch that is to fit the die is squared and coated with layout dye. The
punch contour may be layed out either by scribing around the master template or by
placing the rough punch face against the die opening and scribing around the die
opening. The surplus material is then removed with the contour band saw and die-
filling machine. In pedestal punches, the surplus material is generally removed with
the vertical milling machine and rotary file. The punch is roughed out until it is from
0.13 to 0.38 mm larger than the die opening.
Toolmaking Practises 101

An exact match of the punch and die is obtained by mounting both the punch and die
in the die set and then forcing the soft punch into the hardened die by means of a
hand-operated hydraulic press. The punch is initially forced into the die about 0.4 to
0.8 mm and withdrawn. The outline produced by the die is coated with layout dye, and
the punch contour is hand- or machine-filled until the file begins to remove the layout
dye on the outline. The punch is again mounted in the die set and forced into the die
a little further. This shaving and filling process is repeated until the punch enters the
required distance. Proper clearance is obtained by continued filling and scraping.
The clearance may be hand-stoned on the cutting edges after hardening the punch if
only a small amount of clearance is required (25 to 50 mm). Larger clearance should
be filed before heat treatment, as hand stoning is a tedious job at best. Clearance
between the punch and die may be checked by feeler gauges or shim stock. Final
adjustments in clearance can be made by shearing paper of varying thicknesses and
observing the places where the paper is clean-cut or drawn over. For example, a
section of the cutting edge that cuts a cigarette paper cleanly may indicate that this
section does not have sufficient clearance and should be opened by hand stoning.
A different approach to the shaping of punches is to attach a soft master template
to the face of the punch, as shown in Fig. 2.110. The master can be attached by
screws if there are holes in the punch or by double-faced pressure-sensitive tape. The
punch is rough end-milled as near as possible to the master template. A rotary file
is then mounted in a high-speed drill press or vertical milling machine. The shank of
the rotary file should be the same diameter as the body of the file. In use, the punch

Equal to
clearance
Template

Punch Punch

(a) (b)

Equal to
clearance

Hardened
removable collar

Punch

(c)
Fig. 2.110 Use of master template and rotary file to shape punch
102 Tool Design

is gently brought to bear against the rotating file until enough material is removed to
allow the template to touch the straight shank of the file. The punch is guided around
the file, with the shank acting as a guide and stop. Sharp fillets cannot be reached by
the rotary file and have to be shaped by hand filling.
Clearance can be provided by making the shank of the file smaller in diameter than
the body by an amount equal to the clearance (see Fig. 2.110b). Some toolmakers
grind the shank of the rotary file undersize and use collars of different diameters to
obtain the desired offset between the file body and shank, as shown in Fig. 2.110c.
Figure 2.111 shows a commercial toolmaking machine made especially for finishing
punches and dies. It dramatically reduces the time to perform die-making operations
which the skilled toolmaker conventionally accomplished by hand or with specialised
equipment. The techniques, which are easy to master, do not require any long retrain-
ing programs. They will broaden the capabilities of the toolmaker and increase his
productive time.
After completing the preliminary rough-mill-
ing and sawing operations, the toolmaker
can completely finish punches, dies, and
stripper plates (including all sharp square
inside corners) with this machine. This elimi-
nates the need for transferring work from
machine to machine and the delays caused
by waiting for specialised equipment to be
released for use.
The toolmaking machine shown in Fig.
2.111 basically consists of a motor-driven
head mounted above a vernier tilt table. The
motor-driven head is rheostatically controlled
with speeds of 0 to 45,000 rpm. The head
reciprocates up and down, and the length of
stroke is adjustable. The cutting tools used
are generally carbide burrs ranging from 1.5
to 6.5 mm in diameter; however, diamond
points and diamond “shape” tools are also Fig. 2.111 The Roto Recipro toolmak-
available. A stylus is mounted in the table. ing machine (The Producto
Changing the stylus-to-burr ratio controls the Machine Company)
workpiece size.
Figure 2.112 shows the basic steps in shaping a punch and die. In conventional
diemaking, the usual first step is to make a trial piece part. It may be made by machin-
ing to a layout line, as shown in Fig. 2.112a. Once the piece part has been proved, it
serves as a sizing guide or template, as shown at b. The piece part is mounted on the
bottom of the rough punch, which has been prepared by rough milling. The piece part
serves as a sizing guide, while the punch is hand-guided along its periphery against
the cutting tool. Changing stylus-to-cutter ration controls the punch size to allow for
clearance.
When the punch is completed, it is heat-treated and coated with a wax-base parting
compound. It is then positioned against a rough band-sawed die block, as shown in c.
Epoxy is troweled around the punch and allowed to set. The punch is then removed
and the epoxy serves as the guide for the machine stylus.
Toolmaking Practises 103

Trial piece parts, trial stampings, and prototypes can be made on the toolmaking
machine, as shown in Fig. 2.112d.

Trial piece parts, trial stampings, and prototypes

Fig. 2.112 Using the Roto Recipro toolmaking machine (a) In simple terms, it is the machin-
ing of the piece part to a layout line (no template need be made) (b) Using the
piece part for a sizing guide to finish-size a punch (c) An impression made with
the finished and hardened punch in a rough band- sawed die block or letting
epoxy set around the perimeter of the punch provides the guiding and sizing
reference for the die opening (d) With the tool reciprocating and rotating, use
the stylus machining technique to make trial parts for developing. Sketch at left
illustrates principle for size control by changing stylus-to-burr ratio. Up to 50
prototypes or stacked stamping samples 1/32 in. thick have been made in one
stack. (The Producto Machine Company)

2.9 ELECTRO DISCHARGE MACHINING


Electro Discharge Machining (EDM) is a relatively new process that is being widely
applied to the manufacture of dies and moulds. It is also used for production machin-
ing of difficult-to-machine parts, but the discussion here will be limited to dies and
moulds.
104 Tool Design

EDM is a process of metal removal by means of electric discharges between a shaped


electrode (cutting tool) and an electroconductive workpiece in the presence of a liquid
dielectric (see Fig. 2.113). In simpler terms, an electric spark is used to erode the
workpiece, which takes a shape opposite that of the electrode (cutting tool). There
are thousands of sparks each second, and each spark produces a crater on both the
workpiece and the electrode, the smaller crater normally occurring on the electrode.
It is believed that the crater is caused by the melting and vaporisation of minute par-
ticles of metal. Chips therefore are produced in the form of hollow spheres, which are
ejected into the dielectric fluid and cooled and then flushed from the gap between the
tool and the workpiece.
In use, the workpiece is mounted to the table of the EDM machine, and the tool is fed
down vertically. The tool never touches the workpieces, and a gap of approximately
25 mm is maintained between the electrode and the workpiece by a servomechanism.
The gap distance can be varied to suit the machining conditions. A direct current of
low voltage and high amperage is delivered to the electrode at the rate of approxi-
mately 20,000 electric impluses per second as the sparks jump from the electrode
to the workpiece. Their effect is to erode a cavity into the workpiece that is the exact
image of the electrode.

Electrode
Workpiece
Dielectric
Power liquid
supply

Fig. 2.113 Basic principle of electrodischarge machine

It is very important that the chips be washed away from the gap area. Stray chips
can cause further sparking of the tool, resulting in unwanted taper and tool wear. The
dielectric fluid is continuously pumped through a filter to separate the chips from the
fluid.
The cutting speed, or metal-removal rate, depends upon the volume of metal removed
by each spark and the frequency with which the sparks occur. The volume of metal
removed per spark can be increased by increasing the current to the electrode. The
work finish, the electrode material, and the workpiece material also affect the metal-
removal rates.
It must be remembered that there is always electrode wear during the EDM process,
and therefore the tool will gradually lose its shape during the machining process.
Tool wear is compensated for by changing electrodes as the machining operation
progresses. It may be necessary to use as many as five electrodes to finish a detailed
mould cavity to the required shape and tolerance. Through-hole operations can be
machined with stepped electrodes, which have the effect of changing to new elec-
trodes. Many EDM machines have a “no-wear” features which may be applied to
finishing operations.
Toolmaking Practises 105

The dielectric fluid is generally a light lubricating oil. The most common electrode
material is graphite, although tungsten, carbide, copper, brass, and zinc alloys are
also used. Electrode materials are selected on the basis of wearing qualities when
applied to specific machining operations.
Figure 2.114 shows a typical EDM machine. The major elements are the power sup-
ply, the servo unit, and the machine table. The power supply provides the current for
the machining operation, while the servo unit supports the electrode and maintains
the spark gap. The machine table provides for adjusting and positioning the workpiece
and also contains the dielectric to maintain the level of the liquid dielectric above the
workpiece.
The manufacture of stamping dies with EDM. The EDM process adapts well to the
manufacture of stamping dies because of the ease with which irregularly shaped
through holes can be machined. Complex holes and odd shapes can be machined
without sectionalising the die. Dies can also be machined after heat treatment, and
therefore warpage during heat treatment is eliminated. Dies made of difficult-to-
machine materials, such as carbide, can be machined at a cost as low as that for steel
dies. Complex dies can be made with less skill because the die opening is machined
from the punches themselves. This also results is precise and uniform clearances. All
these advantages add up to fewer man-hours to make a high-quality die; this means
that tool and die shops can bid more competitively and enjoy larger profit margins.

Fig. 2.114 Typical EDM machine used for diesinking and diemaking applications
(Cincinnati Milacron, Inc.)

The EDM process is not without its problems when used to produce dies. At first, it
was noticed that dies produced by EDM had a marked tendency to fail by cracking
and chipping. Further investigation showed that a shallow white surface layer of metal
could be observed under a microscope. This condition was most evident on tool steels
106 Tool Design

hardened before EDM, and the layer was identified as a rehardened zone produced
by the heat generated by the electric discharges. The surface did not show on tool
steels cut by EDM in the annealed state and then heat-treated. It was decided that the
layer did develop on annealed material but that the subsequent heat treatment had a
tendency to mask the layer.
The white rehardened zone, which varies from 25 to 130 mm, produces a highly
stressed untempered zone on the surface of the tool. To compound the problem, the
highly stressed zone lies on the working surface of the tool. This condition can easily
lead to premature failure of the cutting edge, especially when the rehardened zone is
quite deep.
The rehardened zone can be reduced to a negligible minimum by the proper choice
of tool material, dielectric fluid, and good machining practises. For example, experi-
mentation has shown that the shallowest layer of rehardened material is produced at
a high frequency and weak spark; therefore, the machine operator should use one or
more light high-frequency spark finishing passes to remove the effect of heavy rough-
ing cuts. When a very shallow rehardened zone could be detrimental, as in the case
of tools subjected to shock loads, the zone can be eliminated by additional finishing
techniques, such as grinding, lapping, chemical milling, or electropolishing. Finally,
the tool should be retempered to minimise the stress level. The tempering tempera-
ture should be the highest permitted without reducing the hardness.
The most common method of producing dies by EDM is to mount the die set directly in
the EDM machine and use the punches (mounted on the punch shoe) to machine the
stripper plate and die block. In order words, the punch holder and punch become that
electrode, and the die shoe and die block become the workpiece. Since the electrode
and workpiece must be insulated from each other in order to have opposite polarity
to initiate sparking, a special die set must be ordered. Die sets of this type can be
ordered from the manufacturer with insulated guide pins or with insulated bushings.
Other than this insulation, the die set is the same as a standard conventional die set.
If a ball-bushing die set is used, the ball bushing can be removed after setup.
The punch cannot be used directly as the electrode since there is always some wear
on the electrode. It is therefore necessary to face the punch with carbon or other elec-
trode material before machining the openings in the stripper and die. Various methods
have been used in making tipped punches. One is first to rough out the punch and
then heat-treat it to the required hardness. The electrode material is rough-machined
to the configuration of the punch face and fastened to the punch by cementing, solder-
ing, or mechanical fastening. The punch and electrode assembly is then finishground
to size.
Another method utilises a finished punch that is ground to size and hardened.
Standard stock punches can therefore be used. The electrode blank is attached to the
finished punch and then finished to the punch contour. Toolmaking-machine methods
of diemaking, explained in Section 2.8, work well for this operation. The punch contour
can be used as the template can be made.
The most satisfactory method of attaching the electrode blank to the punch is the
mechanical joint, providing the size of the punch permits the use of pins and screws.
Dowel holes, mounting holes, or stripper holes can be utilised. This type joint will give
the least electric resistivity between the punch and the electrode material.
Toolmaking Practises 107

Solder can be used when tipping punches with metallic electrode material such
as brass, copper, tungsten, etc. Carbon materials can be soldered if they are first
copper-plated. Soft soldering is commonly used in the tipping operation. Very seldom
is hard soldering or silver soldering used unless the temperature involved is below the
drawing temperature of the punch. The soft solder commonly used is 50/50 or 60/40
solder, which melts at 188 to 193°C. Care must be taken not to exceed the tempering
temperature of the punch (around 204°C). Also, after the tip is attached, care must
be taken not to exceed the melting temperature of the solder during conventional
machining or the EDM operation.
The adhesive method is probably the most widely used. The growing use of carbon
and the danger of exceeding the drawing temperature of the punch may dictate the
use of an adhesive. If a nonconductive adhesive is used, care must be taken to ensure
electric conductivity between the punch and electrode. This is accomplished by mak-
ing sure that high points of the surfaces touch each other. This means that punch and
electrode material should be clamped in a vise or fixture immediately after the two
parts are joined.
The adhesive materials recommended by users are Hysol Kit no. 0017 from the Hysol
Corporation, Olean, NY; Eastman no. 910 from the Eastman Chemical Products,
Inc., Kingspoint, Tennessee, USA; and Eccono bond solder nos. 58C and 57C from
Emerson and Cummins, Inc., Canton, Massachusetts, USA. The manufacturer’s
instructions for use should be followed closely for best results.
The total length of electrode material should be 9.6 mm more than the combined total
of the die thickness and the lower shoe thickness, as shown in Fig. 2.115. The elec-
trode material must be long enough to extend into the dummy plate and still not allow
the punch to touch the die.

Fig. 2.115 Electrode set up for finishing, piercing, and blanking dies with EDM
(Cincinnati Milacron, Inc.)

As explained earlier, the punch and electrode are as one unit and must be finished as
such. If the required die clearance is in the range of 12 to 100 mm per side, the elec-
trode and punch have exactly the same cross-sectional dimensions. If the required
clearance is more or less than this, the electrode size will change accordingly because
die clearance is produced by the spark overcut. In other words, the minimum overcut
108 Tool Design

per side is 12 mm and this may be increased to 100 mm per side by the machine opera-
tor’s selection of machining power settings. The power supply can be set to maintain
a predetermined clearance that is uniform around the entire contour of the die. When
the required clearance falls outside the 12 to 100 mm per side range, the electrode
section of the punch must be increased or decreased accordingly.
A relief must be machined on the electrode, as shown in Fig. 2.116. The length of the
relief will be the die thickness plus 3.18 mm. This relief need only be deep enough to
prevent this section from machining but should not be deep enough to cause weak-
ness of the electrode. In general practise, this can be removed by any means, includ-
ing snag grinding.
The completed punch units are now ready to be mounted to the punch shoe.
Conventional mounting methods are used, as explained in Chapter 9.
The die block and stripper plates must be keyed together with pins so that correct
alignment can be obtained with respect to the punch at the time of assembly. Opening
in the die block should be rough-machined larger than the corresponding relief on the
electrode, as shown in Figs 2.116 and 2.117. This can be done by drilling and band
sawing. As indicated in Fig. 2.17, the stock left for removal by the upstroke of the
electrode must be the least 0.80 mm all around the perimeter of the opening. The die
block and stripper plate can be fastened to the lower shoe and coated with bluing. The
die-set halves are then fitted together, and the shapes of the electrodes are scribed
around the die block.
A minimum of 0.8 mm
to 1.6 mm metal remaining
Relief area
on electrode to be cleaned out by electrote
Die plate
Punch dim.

Rough machine
Relief die-block opening
Electrode

Fig. 2.116 Relief on EDM electrode Fig. 2.117 Die-block opening is rough-machined
(Cincinnati Milacron, Inc.) by conventional methods (Cincinnati
Milacron, Inc.)

An alternate method, which is faster if the die has many punches, is to assemble the
die set, put it on the EDM machine, and electrically etch the impressions of all the
punches on the die block and stripper plate by machining for a few minutes to impart
the punch image on the plate. The lower shoe must also be laid out or etched so that
clearance for slug dropout can be sawed at the proper place.
A dummy plate must be made large enough to cover all the die openings, as shown in
Fig. 2.118. Its purpose will be explained later. It should be drilled, counterbored, and
fastened to corresponding tapped holes in the bottom of the die shoe.
A dielectric manifold must be used, as it is necessary to pump dielectric fluid through
the spark gap to flush out machined particles. This does not require accurate machin-
ing except for parallelism between the top and bottom surfaces (see Figs 2.119 and
2.120). The placement of the dielectric base fitting should be as high as practical
Toolmaking Practises 109

when using vacuum flow to avoid gases forming in the upper part of the chamber dur-
ing EDM. These gases could ignite, causing the electrodes to be damaged.

Dummy plate
Flow holes

Fig. 2.118 Method of dummy plate to die shoe (Cincinnati Milacron, Inc.)

Manifold

Vacuum
line

Fig. 2.119 Dielectric manifold for pumping dielectric through spark gap (Cinnati Milacron, Inc.)

Manifold

Fig. 2.120 Placement of dielectronic manifold on die set (Cincinnati Milacron, Inc.)

At this point the die is ready for machining by EDM. The die set is assembled and
installed on the EDM machine, as shown in Fig. 2.120. Note that the die and stripper
are not mounted at this stage. Before the EDM operation is begun, the electrodes
must be broken loose from the punches and placed in their respective cavities. Since
the electrodes are not fastened to the punch, the die block can be mounted on the
lower show above the electrodes (see Fig. 2.121).
After this is done, adhesive is once again applied to the electrode and the punches
again joined. After checking for continuity between the two, the machining operation
is resumed but in an upward direction, as shown in Fig. 2.121. By machining in the
upward direction, the natural taper effect of EDM can be put to use, as it will afford
angular clearance for slug dropout. With machine settings constant, the amount
of overcut on the entry side will be approximately 1.5 times that on the exit side.
Referring to Fig. 2.122, it is readily seen that if the desired land clearance of 43 mm is
obtained, the bottom clearance will be approximately 1.5 times this amount.
110 Tool Design

Die block
EDM from
botom to top

Fig. 2.121 Die block is machined from bottom to top (Cincinnati Milacron, Inc.)

The normal angular clearance for slug


dropout in conventional diemaking prac-
tises usually runs from 0.5 to 1° (see
Chap. 9). The difference in overcut,
since correct clearance at the land is
43 mm, would be 0.20 mm if the settings
were held constant. On a 12 mm thick
die block, this would mean the angular
clearance would be seven minutes,
which would be inadequate. If the die- Fig. 2.122 Natural taper effect of EDN
block thickness were 25 mm the angle provides angular clearance when
would be three minutes. machining from bottom to top
(Cincinnati Milacron, Inc.)
Fortunately, it is possible to vary the
overcut with EDM to correct this situation. If the machine is run on the relief portion
at maximum amperage and the power cut back on the land portion, the result is as
shown in Fig. 2.123.

Fig. 2.123 Angular clearance can be varied by machining a relief portion at maximum am-
perage and cutting back power on the land portion (Cincinnati Milacron, Inc.)

After the die block has been machined, it is removed from the lower die shoe. The
stripper plate is mounted on the lower shoe for machining. The servo reverse switch
on the EDM machine is changed from machine-up to the machine-down position.
Toolmaking Practises 111

Since the stripper will be machined from the top down and a bell-mouthed condition
at the top is desirable, the flow of the dielectric fluid is reversed (pressure flow in place
of vacuum flow). The degree of accuracy required in the stripper is not as critical as on
the die plate; therefore, the stripper can be machined at the maximum power setting
obtainable on the EDM machine.
Subsequent machining should be done on the stripper around the opening, as shown
in Fig. 2.124. This allows the stroke of punches to be lowered after the die has been
resharpened. Otherwise, the stripper would interfere with the punch radius.

Fig. 2.124 Method of relieving stripper plate to allow punches to be lowered after
resharpening (Cincinnati Milacron, Inc.)

Upon completion of the stripper, the electrode material can be removed from the
punches by giving the electrodes a sharp tap or, on large areas, by heating the punch
to 300°. The heat causes the adhesive to weaken to the point where the electrode can
simply be lifted off.
A clean-up grind can be taken on the punch faces to remove any carbon or adhesive
adhering to them. The die set can now be assembled and placed in production.
Top-to-bottom machining can be done
Vacuum flow
with wafer-type electrodes, as shown
in Fig. 2.125. This is a twist of the pre-
vious method of machining. A relief is
machined on the electrodes, and the
power is increased after the top of the
cutting length of electrode has passed
Fig. 2.125 Top-to-bottom machining with
the die land. The power should be
wafer-type electrodes (Cincinnati
increased gradually to avoid shorting
Milacron, Inc.)
the electrode. If the electrode does back
out of the cut, it is possible to resume rough machining on the land area. Clean oil and
vacuum flow will minimise this possibility. The electrode must be within 5 mm of the
work to start sparking with no particles in the gap.
A variation of the EDM process, known as upside-down machining, can be used in the
construction of blanking dies. As the name implies, all EDM machining is done on the
die and stripper upside down. In addition to making the die block and stripper open-
ings, the punch holder is also made by EDM. All three of these components must have
common locating holes for positioning on a manifolding fixture, as well as locating at
the time of die assembly.
The punch should be made straight, with no heel or pedestal. Electrode material will
be sized with the punch since the components are to be machined upside down, the
end of the punch which will eventually be mounted (not the end which does the shear-
ing) must be tipped with electrode material. If the end which does the shearing were
tipped, a mirror image of the desired shape would be produced in the die.
112 Tool Design

Figure 2.126 illustrates this technique. After the electrode has machined the die and
stripper, the electrode is removed from the punch. The punch holder is then put on
the manifold and machined by the punch itself. A perfect match between the cavity
and the punch results. The punch is now removed from the dummy plate and fastened
mechanically to the punch holder.

Dummy Plate

Punch

Electrode
Machined by
electrode

Die plate

Stripper
Machined by punch
Punch holder

Pressure
flow
Manifold
fixture

Fig. 2.126 Upside-down EDM machining (Cincinnati Milacron, Inc.)

The advantage of this method is that punches can be kept extremely short, and thus
more rigid, thereby increasing die life. In addition, there is no heel or pedestal on
the punch, and consequently no radius, so that the stripper need not be machined
to accommodate this radius. Figure 2.127 shows how a die machined by this tech-
nique would be assembled (turn Fig. 2.127 upside down). These components can be
mounted on die shoes with little trouble since they have common locating holes.

Die block

Punch Stripper
Punch
holder

Fig. 2.127 Assembled die after upside-down EDM machining (Cincinnati Milacron, Inc.)
Toolmaking Practises 113

2.10 ELECTRO DISCHARGE MACHINING FOR


CAVITY APPLICATIONS
There are several advantages to the machining of mould cavities by EDM, all of which
add up to reduced cost. The major advantage is that EDM simplifies the machining of
complicated shapes especially when there is a great deal of fine detail in the bottom
of the mould. Hand finishing is virtually eliminated compared to tracer milling because
scallops are not produced. Successive dies can be accurately duplicated, and it is
generally not necessary to sectionalise the mould.
The ability of EDM to machine hardened materials makes it possible to machine or
at least finish the mould after it is heat treated. This feature is also valuable is cavity
maintenance, as washed out or worn dies and moulds can be reclaimed.
It is not always economical to sink mould cavities by EDM. Cavities that have straight
sides or flat bottoms with little detail and geometric shapes that can easily be machined
should not be sunk with EDM (see Fig. 2.128). This type of workpiece geometry should
be cut with a rotating tool or grinding wheel providing the material can be machined by
these methods. Conventional machining methods are much faster in this case.
When the machinability or workpiece geometry factors prohibit conventional machin-
ing practises, EDM may solve the problem. A typical shape is shown in Fig. 2.128a.
Here the cavity depth-to-width ratio is quite high and would require a long slender cut-
ter to machine by a conventional rotary cutting tool. The deflection of the tool would
almost prohibit machining by this method. Besides, the wall geometry is rather com-
plex, and the problem is compounded by the fact that the geometry is confined by the
narrow opening. If the cavity could be sunk by a tracer mill, it would be very difficult to
provide a suitable surface finish either by machine-or hand-finishing methods.

(a) (b) (c)

Fig. 2.128 The complexity of the cavity determines the method of machining: (a) EDM only
(b) Conventional machining (c) EDM and conventional machining

Figure 2.28c shows a cavity that could best be machined by a combination of EDM
and rotary-tool application. In this case, the major portion of the cavity is roughed out
by conventional methods and the final intricate detail is sunk by EDM. The advantage
114 Tool Design

of this method is that the majority of the material can be removed rapidly. The mould
can be rough-machined and then heat-treated, followed by EDM finishing.
Although it is relatively easy to impart the mirror image of an electrode face in the
bottom of a mould cavity, it is not always a simple matter to manufacture the elec-
trode. In some cases, it may be just as easy to go ahead and sink the cavity from
the master pattern as it is to make the master pattern, transfer it to an electrode face,
and then sink the mould. However, it is generally easier to make a master pattern
for an electrode as the usually (although not always) involves working with external
rather than internal surfaces. Also, in most cases, two to six electrodes are required
to finish-machine a complex mould. In this case, one master pattern to duplicate all
electrodes is required.
Master patterns for EDM electrodes are constructed by techniques described in
Section 2.11. Once the pattern is constructed, the electrode is duplicated by use of
a pantomill or tracer mill. The material used for the electrode must have the ability to
machine the mould material and resist wear. Graphite is most commonly used for this
application; however, a copper tungsten material is often used for cavities containing
sharp corners, precise detail, and fine finish because tungsten will not flake as badly
as graphite when it is machined to small sections.
Electrodes constructed by machining often require hand finishing to remove the scal-
lops left by the duplicating operation. This can be eliminated by using cast electrodes.
In other words, a casting produced in a master die is used as an electrode.
Various grades of zinc and zinc-tin alloys are a popular castable electrode mate-
rial, although aluminium has been used. The most desirable electrode material is a
combination of zinc and tin, since this combination has relatively low shrinkage and
low melting temperature. The zinc and tin can be purchased separately from local
sources, and the alloy is relatively inexpensive pure zinc in that it can be remelted
and used repeatedly. Any combination of zinc and tin from pure zinc to a proportion
of 50/50 can be used as electrode materials. Pure zinc has the best wear ratio, since
tool wear gets worse as tin is added. However, electrode-casting accuracy is best with
50/50 zinc-tin and becomes worse as the proportion of tin is reduced. Due to these
opposing characteristics, it is a normal practise to vary the combination of zinc and
tin in accordance with cavity accuracy and the economics of wear. Recommended
proportions of zinc and tin for various jobs are given in Table 2.3.
Many tool and die shops have equipment which can be used or adapted for use in
manufacturing zinc-tin die-cast electrodes. Generally, the electrode material is melted
in a pot, poured into a preheated mould or existing die, and then pressed to force the
material into the cavity shape. Equipment for this process includes a thermostatically
controlled melting pot for melting the electrode material, a thermostatically controlled
electric hot plate (gas-flame torches and ovens have been used) to preheat the mas-
ter die cavity (to approximately 200°C), and a hydraulic press (approximately 25 Mpa
pressure is needed with a volume of 3.5 to 4 cm3) for cold coining the electrode mate-
rial in the master-die cavity. Supporting equipment such as ladles and safety devices,
including asbestos gloves and eye protection are also needed.
The master die for casting usually consists of an existing die or master steel cavity. A
framework and a pressure plate are required to tool the die for pouring electrodes and
should be made to suit a particular die. If the master die lends itself to exterior framing
without making the electrode excessively large, the frame is simplified. The framing
Toolmaking Practises 115

plates must be thick enough (13 to 25 mm) to withstand pressing forces and extend
high enough (25 to 50 mm) above the die to give the electrode a stable body as
shank. The framework should also extend approximately 2 mm below the face of the
die. Bolts are satisfactory fastening devices to allow the framework to be assembled
and disassembled.
The pressure plate can be one of two designs: (i) of flat stack (approximately
25 mm thick) or (ii) a plate with an extended boss to impart a dielectric manifold in the
back side of the electrodes and also provide locating surfaces for repetitive electrode
mounting. The pressure plate should clear the inside of the framework by a few thou-
sandths. Larger dies may require up to 0.8 mm clearance.
The zinc-tin die-casting material melts at approximately 338°C. A temperature of
427°C has been found desirable to maintain the thermostatically controlled pot. Actual
pouring temperature will be somewhat less. Small-detail dies should be preheated
to approximately 216°C (177°C on large dies) to prevent premature chilling of the
molten material, which would cause voids and cold shock in the electrode castings.
Preheating can be accomplished with a furnace, a hot plate, or by pouring not zinc-
tin into the cavities three or four times before pouring the actual electrodes. In either
case, facilities should be located conveniently to the press to aid in the handling of
the dies.
The preheated master die with framework in position is placed on the press. The
molten electrode material is poured into the master cavity, filling the deepest area
first. The cavity should be filled to a depth to provide the electrode with a stable body
(19 to 38 mm). The pressure plate is positioned within the framework and pressure
is applied as the material solidifies. The operator should control the pressure at this
point to prevent excessive material from being forced out around the edges of the
pressure plate. After the electrode material has solidified, additional pressure is built
up to make the zinc-tin more dense. Pressure of 3.5 to 20 Mpa have been used,
depending upon the size of the electrode and the details in the cavity.
The number of electrodes per finished cavity vary with the accuracy, detail, and vol-
ume of the cavity. As many as 12 electrodes and as few as three have been used on
jobs. On a production run of a die, the last two electrodes (finishing electrodes) of one
die can be used as the first two roughing electrodes of the next die.
Cold coining is used to compensate for the normal shrinkage characteristics of the
electrode material. The coefficient of expansion per foot for 50/50 zinc-tin from melt-
ing temperature to room temperature (without pressure) is 0.69 mm. Naturally, the
electrode will not shrink this amount when the above method is used; however,
some shrinkage occurs. After the electrodes and master die have cooled to room

Table 2.3 Recommended proportions of zinc and tin for castable electrodes
Type of job Zinc, % Tin, %

Detailed 50 50
Moderate tolerances 70 30
Average detail 80 20
Low tolerance 100 0
SOURCE: Cincinnati Milacron, Inc.
116 Tool Design

temperature, the electrodes are replaced on the press and coined. Cold coining pres-
sures vary, as do casting pressures. Pressures up to 41 Mpa have been used.
The fact that more than one electrode is needed to machine a cavity dictates that
subsequent electrodes duplicate the initial electrode and be located on the machine
platen quite accurately with reference to the partially machined cavity. Subsequent
electrodes must be precisely located in order to provide accuracy and detail in the cav-
ity. There are various locating techniques used in positioning electrodes on the platen
of the EDM machine. Three will be discussed here, in order of their preference.

First technique The first (and most acceptable) way is to internally cast locating sur-
faces into the back side of the electrodes during hot pressing and cold coining. In this
manner, the locating impression of the electrodes can be repeatedly maintained in the
exact relationship with the electrode face contour. At the same time, the relationship
of the back sides of the electrodes with the face contours is uniformly maintained.
Then all that needs to be done to locate subsequent electrodes on the machine is to
locate and secure locating pads on the platen. This technique eliminates the need for
conventional machining of locating surfaces and can also provide a cast-in manifold
necessary for dielectric flow.

Second technique The second method is to machine the back side and two adjacent
edges of the cast electrodes conventionally and locate on the platen against three
pins. The machining operation will determine the accuracy of locating electrodes and
can best be done by placing the electrodes in the master die for the machining opera-
tions. There is no particular reason to maintain uniform thickness on all the electrodes,
since the purpose of machining the back side is to establish a flat reference plane with
the contour face. Of course, the reference plane must be kept in the same relationship
with the contour face on all electrodes.
The adjacent edges of the electrodes must be machined quite accurately in reference
to checkpoints on the contour face. The best approach to this problem is to machine
one edge of all electrodes at one machine setting and then machine the adjacent
edge of all electrodes at another machine setting. In both operations the electrodes
should be located in the master die to establish the edges in reference with check
points on the contour face.

Third technique The third technique of electrode location is a flagging method. The
die block must first be partially machined on two edges and the top side to establish
reference planes. To align the electrode contour face with the top side of the die block
levelling and gauges can be used. An adapter plate is mounted to the machine platen
and used to hold the electrode with levelling screws. The levelling screws, with the
use of gauges, permit adjusting the electrode contour face to the correct relationship
with the die. The gauges are positioned between the die and the electrode at particu-
lar checkpoints, and the levelling screws are adjusted until the gauges make proper
contact with the checkpoint. To locate the electrodes in the correct horizontal position
over the die block, gauge must be made to position against the machined edges of
the die and flag particular surfaces of the electrode contour. This will position the elec-
trode in the correct relationship with the die block in the cross and longitudinal direc-
tions. Either the die block or the electrode can be shifted to obtain this relationship.
The flagging technique permits locating subsequent electrodes without machining
locating surfaces on each electrode. However, this technique does require flagging
each new electrode and would be only as accurate as the operator performing the
operation.
Toolmaking Practises 117

In any of these three locating techniques, actual mounting or securing of the elec-
trodes to the machine platen can be accomplished with toe clamps or hold-down
bolts. If hold-down bolts are used, the electrodes must be uniformly drilled and pos-
sibly tapped.
A machining operation must be performed before the cast electrodes can be used.
This is the drilling of dielectric flow holes. Dielectric fluid must be introduced into the
machining gap during the cutting operation to wash the gap of metal particles, cool
the work and electrode, and help initiate the spark discharge.
The holes are located in critical areas of the electrode contour and are generally
placed where pockets or projections exist. The size of the holes depends upon the
particular application, i.e., forging dies can afford to have larger dielectric flow holes
than plastic moulds. Generally speaking, sizes can vary between 0.82 to 3.2 mm.
The smaller holes are preferred, since slender “cores” are left in the cavity where the
electrode is drilled. Smaller the holes, smaller are the cores.
The dielectric flow holes are generally step-drilled to reduce the thickness of electrode
that must be drilled with a small, short-fluted drill. The necessary holes can be laid out
and centre-punched, both front and back. The back side of the electrode is then drilled
with a 6.4 mm drill to 6.4 mm of breakthrough. The contour face can then be drilled
with a small drill to break into the 6.4 mm hole. The position of these small holes can
be slightly shifted on subsequent electrodes to eliminate the cores produced by the
previous electrodes.
The dielectric oil can be introduced to the electrode through a manifold or individual
plastic lines. In most cases, a manifold is satisfactory and can be readily manufac-
tured for a particular job. Mounting of the manifold should be considered when decid-
ing upon the method of mounting the electrode.
The useful life of zinc-tin electrodes can be considerably increased if the electrodes are
electroplated with chrome. Finer details can also be achieved with plated electrodes.
Electrodes to be plated must be acid-etched to make them thoroughly clear and uni-
formly receptive to the plating action.
Both acid etching and electroplating are techniques that should not be lightly treated.
Each requires a fair amount of skill and application know-how, to say nothing of the
specialised equipment involved. Smaller shops usually do not have the know-how
and equipment available and must send electrodes to custom platers. The following
general information, however, will help the reader familiarise himself/herself with the
process.
For etching steel, see Table 2.4. For etching zinc-tin (50/50) electrodes the etching
solution is made up to 70 per cent water, 10 per cent sulfuric acid and 20 per cent
nitric acid. While etching, it is necessary to remove the electrode every 60 seconds
so that gas is not trapped in pockets of the electrode. Also, one must make sure that
the dowel holes match the edges and that the machined backs are all protected with
Micropeel or the equivalent. Table 2.5 gives the formulation for brass. For etching
tungsten materials, under controlled conditions with the hood exhausted to atmo-
sphere, considerable gas is created including fluorine. The etchant is hydrofluoric
aqua regia, consisting of 1 part nitric acid and 3 to 4 parts of hydrofluoric acid.
118 Tool Design

Table 2.4 Formulation for etching steel


37% hydrochloric acid Water Temp, °C Rate per min, mm per side

80% 20% 57.22 2.5

The solution must be agitated every minute to remove entrapped gases. Etching rate will change
as acid is consumed. Variations in temperature will also affect rates. Mask the unetched portion
with Microstop or Micropeel or equivalent. After etching, neutralise with 50% solution. (Microstop
lacquer and Micropeel are available from Michigan Chrome and Chemical Co., Detroit)

Table 2.5 Formultion for etching brass


70% nitric acid Water Temp, °C Rate per min, mm per side

20% 80% 46.11 5


30% 70% 29.44 5

The recommended procedure for hard chrome plating zinc-tin electrodes is as


follows:
1. The electrode should be vapour-blasted or burnished with steel wool.
2. The electrode should be cleaned in a hot alkaline solution.
3. It should be dipped in a 8% sulfuric acid solution.
4. The sulfuric dip is left on the electrode, and it is placed in the chroming tank.
Caution: If the sulfuric dip is removed by rinsing, oxidation may take place
before the electrode can be placed in the chroming tank.
5. A low voltage (as low as possible) should be applied first, and current is then
increased intervals of 30 second until maximum current is reached.
6. Maximum current should be limited to 157 mA per mm.
7. The application rate of hard chrome on the edges or apex of a cutting surface
will be at the rate of approximately 33 mm hard chrome.
The roughing electrodes are used as cast and coined if programmed amperage
reduction is used but must be etched in any perpendicular areas (Normally etched all
over). Cut must be stopped short of final depth. The finish electrodes are plated with
50 to 130 mm hard chrome.

2.11 TRACER AND DUPLICATING MILLS FOR


CAVITY APPLICATIONS
When using tracer-controlled milling machines for cavity applications, the form of a
master pattern is transferred to the workpiece through a tracer unit activated by a
stylus (tracer finger) that follow the master pattern (see Figs. 2.129 and 2.130). The
workpiece material and master pattern are both clamped to the milling-machine table
so that as the table is fed horizontally or transversely, the master pattern is passed
beneath the stylus. The form of the master pattern causes a rise and fall in the stylus
spindle, which through a hydraulic or electrohydraulic circuit causes a corresponding
rise or fall in the cutter. The resulting workpiece profile is therefore identical to that of
the original master pattern.
Toolmaking Practises 119

Fig. 2.129 Using an automatic electrohydraulic tracer milling machine to produce a


cake-pan mould (Tool Products Company, Inc.)

Fig. 2.130 Completed cake-pan mould (Tool Products Company, Inc.)

Types of tracing equipment The simplest type of tracer control is the 180°, or depth-
control, tracer unit. Here only the vertical machine movement, either the vertical head
or the knee, is under tracer control. The other table movement, longitudinal or trans-
verse, operates at a constant feed rate to move the master under the stylus. Thus, the
tracer controls the movement of the vertical head and milling cutter toward and away
from the workpiece as dictated by the shape of the master. After each pass across the
workpiece, the table must be indexed over to a new position and the feed reversed in
order to take a series of incremental cuts. This method of tracing is often referred to
as scanning. Figure 2.131 shows the cutter path during a scanning operation.
Many tracing machines are equipped with options or accessories that make it pos-
sible to do 180° scanning automatically. Automatic scanning requires the use of a
pick feed, which moves the table over the correct increment for the next cut when the
feed is reversed. In operation, the feed is reversed by trip dogs which have been set
for the maximum distance required to scan the widest part of the workpiece. As the
120 Tool Design

feed is reversed, the machine table automatically pick feeds over the selected pick
increment. The pick increment is the dimensions between the lines of cut and may be
adjusted from 130 to 50 mm depending upon the particular machine.

Automatic
Scan length Scan width
180° Automatic

Fig. 2.131 180° pick-feed scanning (Gorton Machine Corporation)

An important feature of the automatic cycle that is not readily noticed is the control
of the feed rate of the table in relation
to the steepness of the cavity shape.
A uniform cutting feed rate should be
maintained to ensure uniform chip loads
regardless of the shape or slope of the
master pattern. In other words, as the
slope increases, the feed rate automati- (a) (b)
cally is reduced, and as the tracing path 360°
levels out, the feed rate increases.
A second type of tracer control is the
360° profiling tracer unit, in which the
tracer, through the master pattern,
controls the feed rates of horizontal
table movements simultaneously (see (c) (d)
Figure 2.132). This type of tracing per-
mits reproducing shapes in a horizon- Fig. 2.132 360° profiling: (a) ID profile
tal plane through an included angle of (b) OD profile (c) ID step
360°. profile (d) OD step profile
(Gorton Machine Corporation)
Figure 2.133 shows the principle of
operation of a 360° tracing unit for one particular machine. The operator depresses
the slide-control push buttons on the operator’s control panel to move the template
toward the tracer head until the stylus contacts the edge of the template. This deflects
the stylus from its centre position, establishing a deflection line and deflection angle.
On this particular machine, the deflection angle is measured electrically. The deflec-
tion line is at a right angle to the template, and the tracer stylus will move along this
line until the deflection preselected on the deflection control is obtained. As the tracing
stylus moves along the edge of the template, any over or underdeflection will cause
the tracer to move along a line at a right angle to the template until the reference
deflection is reached. This is one of two characteristics of the tracer system control-
ling automatic machine-slide movement.
Toolmaking Practises 121

Stylus path following


template edge

x
90°
x = reference deflection
0.005 to 0.015 preselected
by operator Line of deflection Velocity line
at right angle to tangent to template
template edge

Fig. 2.133 Principle of operation of 360° OD profiling on Gorton Auto-Trace Master


(Gorton Machine Corporation)

The second characteristic of the control system is designed to cause a feed along the
edge of the template, regardless of shape. A line established at right angles to the line
of deflection results in a tangent to the template. This is the line along which the tracer
will move under feed. As the template contour changes, the lines at right angles and
parallel to the template edges change, causing the tracing stylus to be continuously
directed along a course tangent to the template at the contact point.
The direction in which the machine slide moves is automatically changed to corre-
spond to the direction and shape of the template edge (Direction of trace and feed
rate are selected at the control station). These two characteristics provide the basis
for the automatic tracer operation shown in Fig. 2.133.
The ultimate in tracer controls is one that will do both 180° scanning and 360° profil-
ing on one machine. This type of machine is excellent for toolmaking applications,
where there is a constant change of workpiece requirements. One part may be ideal
for profiling, the next for full three-dimensional duplicating, and the next for scanning
pick-feed duplicating.
In three-dimensional duplicating, the table, saddle, and knee move simultaneously or
in combinations or movements in response to the tracer movements over the master.
Three-dimensional control provides the operator with a selection of tracing motions
best suited to the shape and configuration of the workpiece. Figure 2.134 illustrates
122 Tool Design

the machining of a three-dimensional mould section. The selection of either scanning


duplication or profiling permits the area and shape of the mould section to be dupli-
cated in the most suitable manner.

Fig. 2.134 Automatic duplication of three-dimensional mould section


(Gorton Machin Corporation)

When duplicating three-dimensionally, the operator can, through finger pressure,


direct the tracer in radial motions, circular motions, sweeping vertical motions, or
combinations of motions. This type of operator-guided tracing is commonly known
in the metal working industry as pencil trace. The operator guides the tracer stylus
over and around the master pattern in the same manner in which a pencil is held for
writing. Pencil trace is particularly useful in die and mould manufacture to facilitate
scrubbing out pockets remaining after automatically profiling a shape and to duplicate
complex part shapes having many irregular and changing contours.
Figure 2.135 shows an example of pencil tracing. A three-dimensional machine is
used in the duplication of a large two-cavity metal pattern. This type of machine is
extremely versatile in that it has full two- and three-dimensional profiling and contour-
ing capabilities. It is a hydraulic servo tracer miller that is operator-guided with light
finger pressure on the tracer stylus.
An example of how a machine equipped with combination 360° profiling and 180°
duplicating can save time is shown in Fig. 2.136 and 2.137. In Fig. 2.136, a mould
contour is duplicated in rise-and-fall pick-feed mode. In Fig. 2.137, the parting line is
automatically 360° profiled without moving the workpiece. The workpiece is a vacuum-
cleaner die-casting mould section.

Selector of tracer and cutters A light axial deflection of the stylus must take place
before the tracer valve is in position to control the movement of the cutter relative to
the workpiece, as determined by the contour of the master pattern. To compensate
for the initial stylus deflection, it is usually necessary to make the diameter of the
stylus slightly larger than the diameter of the cutter. The toolmaker should refer to the
operator’s manual for instruction for stylus size for a particular make and model of
machine.
Toolmaking Practises 123

Fig. 2.135 Pencil tracing on a three-dimensional profiling and contouring machine


(Gorton Machine Corporation)

The selection of the diameter and shape of the tracer cutter and stylus will be largely
determined by the contour and detail of the workpiece to be milled. The cutter and
stylus must be small enough for the material to be machined away from the smallest
detail. It is often necessary to rough out the workpiece with a large stylus and cutter
and then change to a smaller stylus and cutter to finish the detail. For example, the
cake-pan mould shown in Fig. 2.130 was rough-cut with 25.4 mm high-speed-steel
ball end mill, 9.53 mm penetration, 1.27 to 1.53 pick, and 0.153 m/min feed rate. The
finish cut was with a 9.53 mm carbide ball end mill, 0.508 m/min feed rate, 1.02 to
1.27 mm pick.
Cutters used in diesinking generally have two flutes with a ball end. They may be
straight or tapered. Tapered cutters are stronger than straight cutters of the same size,
especially in smaller sizes. Tapered cutters may be used as long as the die cavity has
sloping sides. The angle of the taper in the cutter and stylus should be somewhat less
than the angle of the sidewalls in the die. Straight ball cutters and stylus must be used
when milling straight sidewalls in a die cavity.
Single cutting-edge cutters with sharp points are sometimes used for working out fine
detail. The cutter point should be rounded to a 0.13 mm radius.
It is sometimes possible for the toolmaker to alter the workpiece size to some extent
by using an undersize or oversize stylus. For example, it is possible to make a male
and female mould that differ in size by the wall thickness of the mould product from a
single master. A female mould is first made from the female master and then a male
cast is made from the female mould. The male cast is then used as a pattern. An
undersize stylus is used to compensate for the oversize in the male cast. The result-
ing male mould will be quite close in final size and shape.

Pantomills Pantomills, like tracer and duplicating mills, produce a workpiece of the
desired shape by tracing a master pattern. However, pantomills are quite different
from tracer and duplicating mills in both appearance and operation. One of the major
differences is that the milling cutter is connected to the stylus through a precision
124 Tool Design

mechanical linkage that can be adjusted to change the size relation between the mas-
ter and workpiece from 1:1 to a reduction as great as 40:1, depending upon the type
of machine. Pantomills generally work from an oversize master to effect a reduction
of any errors or irregularities in the master.

Fig. 2.136 Duplicating a mould contour in 180° scanning pick-feed duplicating


(Gorton Machine Corporation)

Fig. 2.137 Profiling the parting line of the workpiece shown in Fig. 2.136 by 360°
profiling (Gorton Machine Corporation)

Pantomills are available in two-dimensional models, which are extensively used for
precision engraving, or in three-dimensional models, used for die and mould produc-
tion or machining of complex shapes. Figure 2.138 shows a three-dimensional panto-
mill applied to the machining of a turbine-wheel pattern or mould. The master in this
illustration was developed from a smaller part section through the use of the versatile
enlarging attachment on the machine. The Lucite or plastic master developed with
this method is shown in Fig. 2.138 being used in the duplication of the turbine wheel
at a ratio.
Toolmaking Practises 125

Fig. 2.138 Pantomill used to machine turbine-wheel pattern mould


(Gorton machine corporation)

2.12 LOW-MELTING TOOLING MATERIALS


The toolmaker and tool designer are frequently called upon to build or design tooling
for workpieces or products that are not easily adapted to conventional tooling prac-
tises. It is sometimes necessary to stretch the imagination in order to come up with
workable solutions for problems of this nature. Low-melting tooling materials may
solve the problem in many instances.

Low-melting bismuth alloys One of the prominent low-melting tooling materials is


bismuth alloyed with lead, tin, cadmium, indium, or other metals. The result is an alloy
with a very low melting point that expands when it solidifies. Many of these alloys are
liquid at temperatures below the boiling point of water, and a few have melting points
below 150°.
Bismuth in a pure state expands 3.3 per cent of its volume upon solidification, but
when it is alloyed with other metals, the expansion is modified according to the rela-
tive percentages of the alloying elements. This property, along with the low melting
point, makes this material suitable for certain tooling applications. Most metals shrink
when solidified in moulds and tend to pull away from the mould walls. Since bismuth
alloys expand and push into the mould detail when they solidify, they are excellent
for duplication and reproduction purposes, as well as for serving as a matrix to hold
fragile workpiece.
It is not necessary for the toolmaker to be concerned with the percentages of alloying
elements. A number of bismuth alloys suitable for low-melting alloy tooling applica-
tions have been standardised and can be purchased commercially. Tables 2.6, 2.7,
and 2.8 give the physical properties of standard alloys as an aid for selection.
The application of low-melting bismuth alloys to tooling is limited only to the imagi-
nation of the toolmaker. These alloys have been used for punch and diesetting,
anchoring, chucks, cores, moulds, patterns and fixture nests–to name a few. The
126 Tool Design

following discussion describes various applications to familiarise the beginner with


the process.
The simple set of removable bismuth-alloy jaws illustrated in Fig. 2.139 is made to
hold a forged-brass piece during a milling operation. The jaws fit a machined cavity
in a steel or brass holder so that they can be removed and jaws for other shapes
inserted.

Fig. 2.139 Bismuth-alloy jaws made to hold a forged-brass piece during a milling
operation (Cerro Copper and Brass Co.)

Figure 2.140 shows a magnetic chuck made with bismuth alloy. Since bismuth is a dia-
magnetic metal, it has been useful as a means of separating pole pieces on magnetic
chucks. Unlike lead or babbitt, which shrink on solidifying, it does not require peening
to make it fit tightly, for it expands and makes all joints watertight.

Fig. 2.140 Magnetic chuck made with bismuth alloy (Cerro Copper and Brass Co.)

In Fig. 2.140 the four narrow rectangular pieces are permanent Alnico magnets fas-
tened with bismuth alloy. The magnets hold a metal cover on a metal can while it is
being lock-seamed

Melt-away tooling Another low-melting tooling material is marked by M Argueso and


Company under the trade name of Rigidax. Rigidax is a castable, thermoplastic com-
pound for stabilising, supporting, and holding hard-to-fixture, thin-wall, odd-shaped,
and nonmagnetic parts. It is melted to fluid form and poured in or around the part.
After machining, the Rigidax is melted out and can be reused.
Toolmaking Practises 127

Table 2.6 Cumulative growth and shrinkage of low-melting bismuth alloys, Test bar 12.71
× 12.71 × 254.2 mm, weighing approximately 4.45 N
Alloy* Time after casting
Minutes Hours

2 6 30 1 2 5 7 10 24 96 200 500

Eutectics:
Cerrolow† 117 +5 +2 0 –1 –2 –2 –2 –2 –2 –2 –2 –2
Cerrolow† 136 +3 +2 +1 0 –1 –2 –2 –2 –2 –2 –2 –2
Cerrobend† +25 +27 +45 +51 +51 +51 +51 +51 +51 +53 +55 +57
Cerrobase† –8 –11 –10 –8 –4 0 +1 +3 +8 +15 +19 +22
Cerrotru† +7 +7 +6 +6 +6 +5 +5 +5 +5 +5 +5 +5
Noneutectics:
Cerrosafe† +20 +22 +40 +46 +46 +46 +46 +46 +46 +48 + 50 +52
Cerrolow† –4 –7 –9 0 +16 +18 +19 +19 +22 +25 +25 +25
Cerromatrix† +8 +14 +47 +48 +48 +49 +50 +50 +51 +55 +58 +61
Cerrocast† –1 –1 –1 –1 –1 –1 –1 –1 –1 –1 –1 –1
*For composition, see Table 2.7
† Trade mark of Cerro Corporation
SOURCE: Cerro Corporation

Table 2.7 Percentage composition of low-melting bismuth alloys


Element Eutectics Noneutectics
Cerrolow* Cerrolow* Cerro- Cerro- Cerr Cerro- Cerro Cerro Cerro
117 136 bend* base* otru* low* safe* matrix* cast*

Bismuth 44.7 49.0 50.0 55.5 58.0 48.0 42.5 48.0 40.0
Lead 22.6 18.0 26.7 44.5 25.6 37.7 28.5
Tin 8.3 12.0 13.3 42.0 12.8 11.3 14.5 60.0
Cadmium 5.3 10.0 9.6 8.5
Indium 19.1 21.0 4.0
Antimony 9.0
* Trade mark of Cerro Corporation
SOURCE: Cerro Corporation

Rigidax is generally applied in one of the three ways:


1. The part is cast (single or multiple) in Rigidax through the use of a baseplate
and damming box. The cast parts are then machined as required. Following the
machining operations, the Rigidax is melted out of the part.
2. Rigidax is cast in the part. Thus the part is converted to a solid to eliminate
chatter and vibration during machining. This is especially useful where it is
necessary to machine workpieces that contain thin fins or flanges.
3. In thin-film clamping, a film of Rigidax is used to clamp small electronic,
ceramic, plastic, honeycomb-core, and nonmagnetic parts in high-production
operations.
Table 2.8 Properties of low-melting bismuth alloys
Alloys Melting temp, °C Range, °C Yield temp, °C Weight, Specific gravity Tensile strength, Elongation ‡ in 50 BHN
kN/m3 at 20°C MPa mm slow loading, %

Eutectics:
128 Tool Design

Cerrolow* 117 42.2 ... 42.2 2.2 8.9 37 1.5 12


Cerrolow* 136 57.8 ... 57.8 2.1 8.8 43 50 14
Cerrobend* 70 ... 70 2.3 9.4 41 200 9.2
Cerrobase* 123.9 ... 123.9 2.6 10.3 44 60–70 10.2
Cerrotru* 138.3 ... 138.3 2.2 8.7 55 200 22
Noneutectics:
Cerrolow*† 61.1–65 63.9 2.4 9.45 34 13.5 11
Cerrosafe*† 71.1–87.8 72.5 2.3 9.4 37 220 9
Cerromatrix*† 103.3–226.7 115.6 2.4 9.5 90 Less than 1 19
Cerrocast*† 138.3–170 150 2.0 8.2 55 200 22

Coefficient of Specific heat‡ Latent of fusion‡ Thermal conductivity‡ Electric conductivity


thermal expansion‡ kCal/kg (solid), cal/(cm2) based on pure
per °C (0.0000 omitted) Liquid Solid (°C) (sec) [Cu = 0.94] Cu, %

25 0.035 0.035 3.33 ... 3.34


23 0.032 0.032 4.44 ... 2.43
22 0.040 0.040 7.88 0.045 4.17
21 0.042 0.03+ 4.00 0.04 1.75
15 0.045 0.045 11.11 0.05 5.00
22 0.040 0.039 3.89 0.045 3.27
24 0.040 0.040 5.56 0.05 4.27
¶ 0.04 0.045 ... ... 2.57
15 0.047 0.047 11.23 0.09 7.77
* Trade mark of Cerro Corporation
† There is no definite w melting point; see the yield temperature
‡ Approximate
¶ At 15.5°C to 54.4°C = 0.000022, at 54.4°C to 103.3°C = 0
Table 2.8 Properties of low-melting bismuth alloys–continued
Resistivity, ohms Maximum loads MPa Safe sustained load, ‡ MPa Volume change,%
30 sec 5 min Liquid to solid Linear growth after solidification

0.5180 ... ... ... –1.4 Less than 0.05


0.7081 ... ... ... –1.35 Less than 0.05
0.4135 69 28 2 –1.7 0.6
0.8825 558,000 28 2 –1.5 0.3
0.3445 103 62 3 +0.77 0.05
0.5282 ... .... ... – 1.7 0.6
0.4037 62 26 2 –2.0‡ 0.3
0.6696 110 69 2 –1.5‡ 0.5
0.2219 103 65 3 +0.5‡ 0‡
‡ Approximate
Toolmaking Practises 129
130 Tool Design

When Rigidax is used for production tooling, it is well to use a handling system
designed to fit the needs of the user. Improper handling can increase the cost of
Rigidax to the extent that its use is prohibitive. Basically, a Rigidax handling system
includes the equipment that will perform the following operations for economical and
efficient handling of the compound:
1. Preheating the part and fixture
2. Melting and dispensing Rigidax
3. Solidifying
4. Melting out and reclaiming
The tool designer or toolmaker should remember that the Rigidax, like other low-
melting tooling materials, is not a cure-all for all tough tooling problems. He should get
in touch with the manufacturer of the tooling material when he has a problem and ask
for his recommendations. His knowledge of the product can often save considerable
time and money. In other words, the designer should check with the manufacturer
before finding by trail-and-error methods that the product will or will not work.

Summary

A designed tool needs to be manufactured. A toolmaker is generally entrusted with


the responsibility of manufacturing the tool, in accordance with specified accuracy
and tolerances. This chapter to familiarises the readers with common toolmaking
practises.
• Machine tools with added features, with high level of accuracy and finish are
used in toolmaking practises.
• Nonconventional machine tools are also from integral part of toolmaking
practises.
• Some of the important tools and machine tools that are used in toolmaking
include: toolmaker’s dial test indicator, sine bars and sine plates, toolmaker’s
buttons, height gauge, toolmakers clamps, edge finder, toomaker’s micro-
scope, grinding machine, boring machine, EDM, etc.

Questions
1. Why is it important for the tool designer to be familiar with the toolroom equip-
ment used by the toolmaker?
2. Why is it important for the tool designer to establish cordial relations with the
toolmaker and other people in the shop?
3. What is the major difference between a dial test indicator and a regular dial
indicator?
4. What is the correct relationship between the swivel contact point and the work-
piece when using a dial test indicator?
5. How is the dial-test-indicator action reversed?
6. What is the major use of the diemaker’s square?
7. What is the advantage of a toolmaker’s clamp compared to a C clamp?
Toolmaking Practises 131

8. What is the purpose of tapped holes in the sides, ends, and top surfaces of
sine plates?
9. When is it advantageous to use a locating microscope?
10. Explain the principle of the continuous-band file.
11. What is the major advantage of the band filling machine?
12. What should be done when the teeth of a band file clog?
13. What is the major advantage of the reciprocating filling machine?
14. How a angular relief angles of dies filed on the reciprocating filling machine?
15. Why are clamps or fixtures not necessary to hold the work on a contour band
sawing machine?
16. How is internal contour saving accomplished on the contour sawing machine?
17. What prevents uneven wear on the grinding wheel of a profile grinder?
18. What two manufactured abrasives are used in abrasive sticks?
19. What is an India oilstone?
20. When is the India stone used?
21. What is the major characteristic of silicon carbide abrasive? When is it used?
22. What is an Arkansas stone?
23. When are Arkanas stones used?
24. How are abrasive sticks or stones prevented from glazing?
25. What are diesinkers riffler files?
26. What is the problem when fine-cut files seem to load too fast?
27. What is the major advantage of pneumatic hand grinders over electrically
driven models?
28. What is a reciprocating handpiece used on flexible-shaft grinders?
29. What materials are used in the manufacture of rotary files?
30. What is the difference between a rotary file and a burr?
31. What determines the safe operating speed for mounted wheels?
32. How are mounted wheels dressed?
33. What are rubberised abrasive points?
34. What is the advantage of rubberised abrasive points?
35. What is the advantage of the square type of abrasive-cloth rotary pad?
36. When are felt bobs used?
37. What type of abrasive, used with felt bobs, is best for fine microfinishing?
38. When polishing with bobs, why is it important to use a series of bobs charged
with successively smaller grain size?
39. Why should threaded tool components that are to be heat-treated be made
from an oil-hard or air-hard tool steel?
40. What should the minimum length of thread engagement when attaching tool
components?
41. Why should a hole that is to be tapped be relieved by countersinking before
tapping?
42. How is heat-treat scale removed from tapped holes?
43. What is a dowel-pin reamer?
132 Tool Design

44. What is the tolerance size of standard commercial dowels?


45. Why should dowel holes extend through both tool components?
46. How are undersize dowel holes finished after heat treatment?
47. Why should the lap always be longer than the hole depth?
48. What machining process is considered the best for producing dowel holes in
hardened components?
49. What is the major disadvantage in using toolmaker’s buttons for precision hole
location?
50. How is final checking done to determine whether a punch is in the correct rela-
tion to a mating die?
51. How does the cutting of soft paper indicate that a punch and die are correctly
aligned?
52. What is transition error?
53. What are the major differences between the vertical-spindle milling machine
and the jig-boring machine?
54. How is the grinding spindle of a jig grinder powered?
55. Why can greater accuracy be obtained with jig grinding compared to jig
boring?
56. What is the disadvantage of converting a vertical milling machine to a jig borer
by adding an accurate measuring system?
57. What is a compensating lead screw?
58. What is the rectangular coordinate system of a jig borer?
59. What is polar-co-ordinate dimensioning?
60. Why is it generally necessary to support workpiece to be jig-bored on
parallels?
61. Why is it important to clamp directly over a supported area of the workpiece?
62. Why are brass heel rests used on jig-borer setups?
63. What two methods are used to align the edge of the workpiece with the
jig-borer table travel?
64. What precautions should be taken to ensure that the table and ways of the
jig borer wear evenly?
65. When it is necessary to reapproach a dial setting, why is it important to back
away from the setting by several thousandths and then make the setting from
the same direction that the original setting was made?
66. How is edge pickup with a slot-type edge finder checked?
67. What is the purpose of the spotting tool shown in Fig. 2.87?
68. What type of tool should be used for rapid enlargement of a hole produced on
the jig borer?
69. When are end reamers used on the jig borer? Why?
70. Why are end reamers held in precision collets rather than drill chucks?
71. What is the advantage and disadvantage of swivel-block single-point boring
chuck?
72. What is the advantage of using the indicating hole gauge shown in
Fig. 2.101?
Toolmaking Practises 133

73. Carefully read and understand the steps given in the discussion of general
operating practises of the jig borer.
74. Why should chips not be allowed to pass through the drill bushing?
75. What is the recommended clearance between the drill bushing and the work-
piece for cast iron and steel?
76. Why are the slip renewable bushings used for reaming longer than the bush-
ings for the drilling operation?
77. What is the result when too much interference is provided on the jig-plate hole
into which a drill bushing is pressed?
78. Read and understand the recommendations given by The American Drill
Bushing Company for the installation of drill bushings.
79. What are file guide?
80. What is the main feature of a rotary cross-slide milling head?
81. What is meant by shearing in a punch?
82. What is electrodischarge machining?
83. Why is it sometimes necessary to use as many as five electrodes when using
EDM to machine and finish a detailed mould cavity?
84. What electrode material is generally used in the EDM process?
85. How is it possible to maintain dies after heat treatment with EDM?
86. Why is it necessary to order a special die set when producing dies by the EDM
process?
87. How is die clearance produced when machining with EDM?
88. When is EDM especially well suited for machining mold cavities?
89. Why is a master pattern to produce EDM electrodes for cavity operations a
necessity?
90. What is the advantage of cast electrode?
91. What material is generally used for cast electrodes? Why?
92. Why is it necessary to apply pressure to a cast electrode as it cools?
93. What machining operation must be performed on cast electrodes before they
can be used?
94. Where are dielectric holes generally located?
95. How can the life of zinc-tin electrode be increased?
96. What is meant by a 180° depth-control tracer?
97. What is a 360° profiling tracer unit?
98. What is a three-dimensional duplicating or tracing machine?
99. Why is it generally necessary to make the diameter of the stylus slightly larger
than the diameter of the cutter?
100. What determines the diameter and shape of the tracer cutter?
101. What is the main difference between a pantomill and a tracer or duplicating
mill?
102. Why do pantomills generally work from an oversize master?
103. What are the unique characteristics of bismuth alloys?
134 Tool Design

References
Books
• Breding, H W: “Tooling Methods and Ideas,” The Industrial Press, New York, 1967.
• Cincinnati Milling Machine Company: “A Treatise on Milling and Milling Machines,” Cincinnati,
Ohio, 1951.
• Connell, R S: “Jig Boring,” The Industrial Press, New York, 1965.
• Glanvil, A B and E N Denton: “Injection-mould Design Fundamentals,” The Industrial Press,
New York, 1965.
• Moore, R F: “Holes, Contours, and Surfaces,” The Moore Special Tool Company, Bridgeport,
Conn., 1963.
• Ostergaard, E: “Basic Diemaking,” McGraw-Hill, New York, 1963.
—: “Advanced Diemaking,” McGraw-Hill, New York, 1967.
TOOLING MATERIALS AND
HEAT TREATMENT
3
3.1 INTRODUCTION
The tool designer is required to select materials for a variety of products, such as
cutting tools, jigs, punches, dies, special machines, and so on. He must have a work-
ing knowledge of these materials and understand their properties and physical treat-
ments. In addition, he must consider the various aspects of tooling, material cost,
fabrication, manufacturing methods, and the proper functioning of the product.
Since the purpose of this chapter is to acquaint the beginning tool designer with tool-
ing materials, it will therefore be limited to these materials. The tool designer should
have a knowledge of other industrial materials, but they are beyond the scope of this
text.
The selection of the proper tooling materials is not a simple undertaking, as there is
no standard set of rules to follow. There are dozens of textbooks, technical publica-
tions, charts, catalogs, etc., dealing with tooling materials, and the tool designer may
translate the information given by these sources to the requirements of his particu-
lar tool-design problem. Many a times, however, this information may not match his
particular job. He is forced to rely on past experience and fundamental knowledge of
tooling materials.
There is a certain amount of trial and error in this approach, and the number of fail-
ures may be quite high, especially when dealing with new tooling materials. The tool
designer should not go overboard with all new materials that appear on the market,
but on the other hand he should not be so conservative that he misses the new and
better materials. The proper course lies between these two extremes. Past experience
and basic knowledge enable the tool designer to make use of the tried and depend-
able tooling materials and at the same time give him a rough idea of the capabilities
of new materials. The discussion in this chapter will help the tool designer acquire
a fundamental knowledge of common tooling materials. Experience can come only
through years of active work as a tool designer.

3.2 PROPERTIES OF MATERIALS


Strength The strength of a material is its ability to resist the action of external loads
or forces. This is the property which prevents the tool from breaking or experiencing
undue distortion under legitimate loads. The loads may be steady, varying, or revers-
ing in definite cycles, and they may be applied in such a manner as to rupture, crush,
or cut the material. Different materials have different abilities to resist different loads
or forces, and there is no simple relationship between various factors.
The tensile strength of a material is the ratio of the maximum tensile load to the
original cross-sectional area of the material. Stated another way, tensile strength is
the amount of smoothly applied pull force per cross-sectional area that a material will
136 Tool Design

withstand before it breaks. Tensile strength is important to the tool designer in the
design of large fixtures, such as those found in the aircraft industry.
The shear strength of a material is the load in Newtons per cross-sectional area
required to produce fracture in the plane of the cross section, the conditions of loading
being such that the directions of force and resistance are parallel and opposite. Shear
strength is important to the tool designer when designing torsion loads, as rupture due
to torsion is of a shearing nature. Knowing the shear strengths of various materials is
also necessary in die design, as they enable the designer to calculate the amount of
force necessary to blank or punch the part.
The compressive strength of a material is the maximum load in Newtons per cross-
sectional area a material can withstand without failure when subjected to compres-
sion. Compressive strength is of importance to the tool designer when using hard-
ened tool steels as a tooling material.
Table 3.1 shows values for tensile, compression, and shear strengths of various
metals.

Plasticity Plasticity is the ability of a material to deform without breaking. The mate-
rial will flow under a load before it ruptures. Some materials will flow to a greater
extent under compressive forces, while others will flow better under tensile forces.
The plastic flow exhibited by a material under compressive forces is known as mal-
leability, while the plastic flow of materials under tensile forces is known as ductility.
Malleability is the property which allows a metal to be rolled into thin sheets, and duc-
tility is the property which allows a metal to be drawn into a smaller-diameter wire.

Elasticity and stiffness Elasticity is the ability of a material to recover its original
size and shape after it has been deformed, and stiffness is the ability of a material to
resist elastic deflection. The modulus of elasticity (Young’s modulus) is a measure of
stiffness, as stiffness is proportional to the modulus of elasticity for identical shapes.
The property of stiffness is important to the tool designer, as the tools he is required
to design usually perform better with the least amount of deflection. This is especially
true when designing tools and parts with long overhang.

Elastic limit The elastic limit of a material is the maximum load per unit area
expressed in N/m2 that can be applied without producing permanent deformation.
Thus, if a metal is acted upon by an external force, it changes shape elastically until
the force reaches the elastic limit of the metal, and then it deforms permanently.

Toughness Toughness is the property of a material that enables it to absorb energy


and deform plastically before failure. Stated another way, toughness is the resistance
of a metal to start permanent deformation plus the resistance the metal has to fail-
ure after permanent deformation starts. A tough material is one that can withstand
heavy loads, suddenly or slowly applied, continuously or often applied, and one that
will deform before failure. There is a definite relationship between the toughness of
a material and its ability to resist shock loads or sudden blows. This characteristic is
called impact strength. The Charpy or Izod notched-bar tests are used to measure the
relative toughness of materials.
Table 3.1 Strength of materials (Stress in MPa)
Metal and alloys Tension, ultimate Elastic limit Compression Bending, Shearing, Modulus of Elongation, %
ultimate ultimate ultimate elasticity × 10–5
Aluminium, cast 103 45 83 ...... 83 758 1–2
Aluminium bars, sheets 165–193 83-97
Aluminium wire, hard 207–448 110-207
Annealed 138–241 97
Aluminium 2–7% 276–345 172
Ni, Cu, Fe, etc.
Aluminium bronze, 517 276 827 ...... ... ... 20
1
5% to 7__% Al.
2
10% Al. 586–689 414
Brass, 17% Zn 225 57 ... 160 .... .... 26.7
23% Zn .... 52 290 3154 .... .... 35.8
30% Zn 194 59 ... 9185 .... .... 20.7
39% Zn 283 120 517 269 ... .... 20.7
50% Zn 214 123 807 5231 ... ... 5.0
Cast, common 124–165 41 207 138 `248 620 22
Brass wire, hard 552
Annealed 345 110 .... .... .... 965
Bronze, 8% Sn 196 131 290 301 .... 689 5.5
13% Sn 203 138 365 238 .... .... 3.3
20% Sn 228 .... 538 391 .... .... 0.04
24% Sn 152 152 786 221
30% Sn 39 39 1013 83
Gunmetal, 9 Cu, 1 Sn 172–379 69 .... ..... 689
Tooling Materials and Heat Treatment 137

Contd...
Contd...
Metal and alloys Tension, ultimate Elastic limit Compression Bending, Shearing, Modulus of Elongation, %
ultimate ultimate ultimate elasticity × 10–5
Manganese,* cast 414 207 862 358
Rolled 689 552
138 Tool Design

Phosphorous ‡ cast 345 165


Wire 689
Silicon, cast, 3% Si 379
5% Si 517
Wire 745
Tobin ++ cast 455
Rolled 552 276 .... ..... ....... 310 35
Cold-rolled 689
Copper, cast 172 41 276 152 202 689
Copper plates, rods, 221–241 69 221
bolts
Copper wire, hard 379–448 ..... .... ..... ...... 1241
Annealed 248 69 .... ..... ..... 1034
Delta meta, ¶ cast 310 .... ..... ..... ...... .... 10
Bars 586
Plates 469 .... .... ..... ..... 1149 17
Wire 689
German silver, 282 130 ..... ...... ..... ..... 28.5
17.2% Zn, 21.1% Ni 138 28 ..... ..... ..... 552 25
Gold, cast
Copper, 5 Au, 1 Cu 345
God wire 207
Iron, cast, common 103–124 41 552 207 124-138 827
Contd...
Metal and alloys Tension, ultimate Elastic limit Compression Bending, Shearing, Modulus of Elongation, %
–5
ultimate ultimate ultimate elasticity × 10

Gray 124–165 .... 172–228


Malleable 186–241 103–138 317 207 276
5
__
Wrought, shapes 331 179 Tensile Tensile Tensile 1930
6
5
__
Bars 345 186 Tensile Tensile Tensile 1930
6
Wire, unannealed 552 ..... ..... ..... ..... 1034
Annealed 414 186 ..... .... ..... 1724
Lead, cast 12 ..... ..... ..... .... 69
Pipe, wire 15–17 ..... ..... ..... .... 69
Rolled sheets 23 ..... 49 .... .... 152
Platinum wire, unannealed 365 ..... .... .... .... 1682 18
Annealed 221 ..... .... .... .... .... 50
Silver, rolled 276
1
__ 3
__
Steel, boiler plates, § firebox 379–448 tensile Tensile Tensile tensile 1999 27.3-23.0
2 4
1
__ 3
__
Flange plates 358–427 tensile Tensile Tensile tensile 1999 28.8-24.2
2 4
3
__
Casting, § soft 414 186 Tensile Tensile tensile 1999 22.0
4
3
__
Medium 483 217 Tensile Tensile tensile 1999 18.0
4
3
Hard 552 248 Tensile Tensile __ tensile 1999 15.0
4
Reinforcing bars, § plain,
3
__
structural grade 379–483 228 Tensile Tensile tensile 1999 25.4–20.0
4
3
__
Tooling Materials and Heat Treatment 139

Intermediate 483–586 276 Tensile Tensile tensile 1999 18.6–15.3


4
Contd...
Metal and alloys Tension, ultimate Elastic limit Compression Bending, Shearing, Modulus of Elongation, %
ultimate ultimate ultimate elasticity × 10–5
3
__
Hard 552 345 Tensile Tensile tensile 1999 15.0
4
140 Tool Design

3
__
Deformed, structural grade 379-483 228 Tensile Tensile tensile 1999 22.7–17.9
4
3
__
Intermediate 483–586 276 Tensile Tensile tensile 1999 16.1–13.2
4
3
__
Hard 552 345 Tensile Tensile tensile 1999 12.5
4
3
__
Cold-twisted ... 379 Tensile Tensile tensile 1999 5.0
4
1
__ 3
__
Rivets, § boilers 310–379 tensile Tensile Tensile tensile 1999 33.3–27.3
2 4
1
__ 3
__
Bridges 317–386 tensile Tensile Tensile tensile 1999 32.6–26.8
2 4
1
__ 3
__
Buildings 317–386 tensile Tensile Tensile tensile 1999 30.4–25.0
2 4
1
__ 3
__
Cars 331–480 tensile Tensile Tensile tensile 1999 30.0–23.0
2 4
1
__ 3
__
Ships 379–448 tensile Tensile Tensile tensile 1999 27.3–23.0
2 4
1
__ 3
__
Shapes, bridges 379–448 tensile Tensile Tensile tensile 1999 27.3–23.0
2 4
1
__ 3
__
Buildings 379–448 tensile Tensile Tensile tensile 1999 25.4–21.5
2 4
1
__ 3
__
Cars 345–448 tensile Tensile Tensile tensile 1999 30.0–23.0
2 4
1
__ 3
__
Locomotives 379–448 tensile Tensile Tensile tensile 1999 27.3–23.0
2 4
Contd...
Metal and alloys Tension, ultimate Elastic limit Compression Bending, Shearing, Modulus of Elongation, %
ultimate ultimate ultimate elasticity × 10–5
1
__ 3
__
Ships 450–467 tensile Tensile Tensile tensile 1999 25.9–22.1
2 4
Steel alloys, nickel, § 3.25%
3
__
N, Shapes, plates bars 586–689 345 Tensile Tensile tensile 1999 17.6-15.0
4
3
__
Rivets 483–552 310 Tensile Tensile tensile 1999 21.4-18.8
4
3
__
I bars, unannealed 655–758 379 Tensile Tensile tensile 1999 15.8-13.6
4
3
__
Annealed 620–724 358 Tensile Tensile tensile 1999 20.0
4
3
__
Copper, 0.50% Cu 414–467 255-262 Tensile Tensile tensile 1999 29.0-23.0
4
Steel springs, 448–758 276-483
untempered
Steel wire, 827 414
unannealed
Annealed 552 276
Bridge cable 1379 655
Tin, cast 24–32 10–12 41 28 .... 276
Antimony, 10 Sn, 1 Sb 76
Zinc, cast 28–41 28 24 48 .... 896
Rolled sheets 48–110
* 10% Sn, 2% Mn
† 9% Sn, 1% P
1 1
‡ 38% Zn, 1__ % Sn, __% Pb
2 3
¶ 55-60% Cu,, 38-40% Zn, 2-4% Fe, 1-2% Sn
Tooling Materials and Heat Treatment 141

§ See specifications of the American Society for Testing Materials


SORUCE: The Watson-Stillman Co.
142 Tool Design

Hardness The ability of a material to resist penetration, scratching, abrasion, or


cutting is known as hardness. Hardness is the property of a metal which gives it the
ability to resist being permanently deformed when a load is applied. Generally speak-
ing, greater the hardness, more will be the resistance to deformation. In the design
of cutting tools, the term red hardness is used to denote ability of the tool to cut at
speeds and feeds that will heat the cutting edge to a temperature approaching red
heat (540°C).
The most widely used method of determining hardness in toolmaking is the Rockwell
test, although the Brinell test may be used in hardness testing of soft materials or
soft materials with a rough surface finish. Rockwell values can be converted to Brinell
values by using Table 3.2.

Table 3.2 Approximate steel-hardness conversion numbers based on rockwell


hardness
Rockwell Brinell
C scale, B scale, Hardness Diam. Vickers Shore Tensile
150 kg, 100 kg, no. 3000 kg sclero- strength,
120° cone 1
___ in. ball 10-mm ball scope MPa
16

68 .... .... .... 940 97


67 .... .... .... 900 95
66 .... .... .... 865 92
65 .... 739 .... 832 91
64 .... 722 2.28 800 88
63 .... 705 2.31 772 87
62 .... 688 2.33 746 85
61 .... 670 2.36 746 83
60 .... 654 2.40 697 81 ....
59 .... 634 2.43 674 80 2247
58 .... 615 2.47 653 78 2172
57 .... 595 2.51 633 76 2103
56 .... 577 2.55 613 75 2034
55 .... 560 2.58 595 74 1979
54 .... 543 2.63 577 72 1917
53 .... 525 2.67 560 71 1855
52 .... 512 2.71 544 69 1806
51 .... 496 2.75 528 98 1744
50 .... 481 2.79 513 67 1689
49 .... 469 2.83 498 66 1648
48 .... 455 2.87 484 64 1599
47 .... 443 2.91 471 63 1551
46 .... 432 2.94 458 62 1510
45 .... 421 2.98 446 60 1462
44 .... 409 3.02 434 58 1420
43 .... 400 3.05 423 57 1386
42 .... 390 3.09 412 56 1351
41 .... 381 3.12 402 55 1317
Contd...
Tooling Materials and Heat Treatment 143

Contd...

Rockwell Brinell
C scale, B scale, Hardness Diam. Vickers Shore Tensile
150 kg, 100 kg, no. 3000 kg sclero- strength,
120° cone 1
___ 10-mm ball scope MPa
in. ball
16
40 .... 371 3.16 392 54 1282
39 .... 362 3.19 382 52 1248
38 .... 353 3.24 372 51 1213
37 .... 344 3.28 363 50 1186
36 109* 336 3.32 354 49 1158
35 .... 327 3.37 345 48 1124
34 108* 319 3.41 336 47 1096
33 .... 3011 3.45 327 46 1062
32 107* 301 3.51 318 44 1034
31 106* 294 3.54 310 43 1007
30 ..... 286 3.59 302 42 979
29 .... 279 3.63 294 41 951
28 104* 271 3.69 286 41 924
27 103* 264 3.74 279 40 903
26 ..... 258 3.78 272 38 876
25 ..... 253 3.81 266 38 855
24 101* 247 3.84 260 37 834
23 100 243 3.88 254 36 814
22 99 237 3.93 248 35 793
21 ..... 231 3.98 243 35 779
20 98 226 4.02 238 34 758
18* 97 219 4.09 230 33 731
16* 95 212 4.15 222 32 703
14* 94 203 4.24 213 31 676
12* 92 194 4.33 204 29 648
10* 91 187 4.42 196 28 620
8* 90 179 4.51 188 27 600
6* 87 171 4.58 180 26 579
4* 85 165 4.67 173 25 552
2* 83 158 4.77 166 24 531
0* 82 152 4.84 160 24 517
*Values are beyond normal range; given for information only
SOURCE: “Ryerson Stock List,” Joseph T. Ryerson and son, Inc., by permission

Machinability The machinability of a material is its adaptability to cutting.


Machinability depends on the hardness and toughness of a material, as well as the
cutting speed, the kind and quality of tool used, cutting-tool angles, the kind and
quality of coolant, the rigidity of the machine, and the rigidity of the tool support. The
standard for machinability is B1112 screw stock rated at 100 per cent. In making a
test, cutting tools of uniform shape, analysis, and heat treatment are used, and com-
parisons are made between the machinability of B1112 screw stock and the tested
material.
144 Tool Design

Since the mechanical properties of tool steel make it difficult to machine, probably at
best the machinability rating would be 30 per cent that of B1112 screw stock. This low
rating requires that the machinability of tool steels be based on W1 water-hardening
tool steel at an arbitrary rating of 100. Table 3.3 gives machinability ratings of tool
steel based on this system.

Table 3.3 Machinability ratings of tool steels water-hardening grades rated at 100
Tool-steel group Machinability rating

W 100
S 85
O 90
A 85
D 40–50
H (Cr) 75
H (W or Mo) 50–60
T 40–55
M 45–60
L 90
F 75
P 75–100
SOURCE: (ASM, “Metals Handbook,” 8th ed., vol. 1, 1961, by permission.

Endurance limit Many tool parts, such as punches and milling cutters, are sub-
jected to a loading cycle in which the stress is not steady but is varying or reversing.
If a large number of cycles are endured during the life of the tool, it may fail although
the greatest stress imposed may be much lower than the yield point of the material.
This type of failure, in which a progressive fracture such as a crack spreads under
repeated cycles of stress, is known as fatigue. Resistance to fatigue depends largely
on the shape of the part, the condition of the surface, and the presence of salt water,
acids, or other corrosive elements. Abrupt changes of section, notches, holes, sharp
corners, and shoulders in a machine part all reduce its ability to resist fatigue, as do
toolmarks, scratches, quenching cracks, and similar defects. But if all these “stress
raisers” are avoided and a specimen of the material is subjected to a complete rever-
sal of stress for more than 107 cycles without failure, the maximum stress endured is
called the endurance limit.

Cost Although cost is not a property of materials, it must be considered in the selec-
tion of material. Prices are constantly changing, and the relative price or availability
of different materials varies from day to day. It is an important part of the designer’s
job to keep up to date in order to make wise choices at the start of a design. Last-
minute changes and corrections cost the company money, and while the designer is
not the final authority on costs, he can do much towards the smooth processing of an
engineering project by making wise choices. Each choice should be made on then
merits of each case, and careful consideration should be given to all the properties
and requirements. Rigid rules and formulas have no place here. In some cases, the
cheapest material may prove most expensive in the long run, and in others the perfor-
mance of an expansive material may well justify the extra expenditure.
Tooling Materials and Heat Treatment 145

3.3 FERROUS TOOLING MATERIALS


3.3.1 Tool Steels
Technically any steel used as a component part of a tool could be called a tool steel.
This is a vague definition, as almost any steel could be used for certain parts of a tool.
To be more specific, a steel containing alloying elements that enable it be heat treated
to obtain desirable characteristics such as strength, hardness, toughness, and wear
resistance could be referred to as a tool steel. This definition is still rather general, as
several grades of standard alloy constructional steel are capable of being heat treated
to obtain the above characteristics.
The term tool steel as used in present-day industry refers to a group of high-quality,
carefully manufactured steels that are characterised by high hardness and resistance
to abrasion. Certain groups of tool steels also have a high resistance to softening at
high temperatures. They are separated from alloyed and heat treatable constructional
steels by precise production and quality-control methods. Tool steels are generally
melted in small-tonnage electric furnaces to ensure the absence of contamination
and obtain more precise control of melting conditions. Tool steels for critical applica-
tions are produced by a vacuum melting process. Rigourous inspection procedures
are employed throughout the manufacture of tool steel. For example, entire bars may
be subjected to magnetic-particle and ultrasonic inspection for surface and internal
defects.
The high price of alloying elements and precise production requirements and qual-
ity control are the reasons for the high cost of tool steels. However, when we stop to
consider that tool steels are made into a complicated tool or die which has required
hundreds of man hours in its manufacture and will produce thousands or even millions
of parts, it is well worth the extra cost.
In the past, tool steels have been classified in several ways. One common method
has been to classify them according to the method of quenching, such as water-
hardening steels, oil-hardening steels, and air-hardening steels. Another method is
to classify them according to alloys, such as carbon tool steels and alloy tool steels.
Still another method has been to classify them according to applications such shock-
resisting steels, cold-work steels, hot-work steels, die-casting die steels, tool and die
steels, etc.
The American Iron and Steel Institute (AISI) and the Society of Automotive Engineers
(SAE) have developed a system of classifying tool steel which groups grades of simi-
lar properties as shown in Table 3.4. This system of classification will be used through-
out this book.

Water-hardening tool steels: group W The water-hardening or carbon steels (group


W) are one of the oldest types of tool steels. They depend primarily upon carbon con-
tent for their heat-treatable properties, with additions of chromium and vanadium.
Chromium is added to increase hardenability and wear resistance, while vanadium
is added to refine the grain for added toughness. Carbon content varies from 0.60 to
1.40 per cent, with approximately 1.0 per cent carbon being most common.
Group W steels are shallow hardening except for small tools under 12 mm in diam-
eter. They will harden with a hard case and tough core and have low resistance to
Table 3.4 Classification and composition of principal types of tool steels
Designation C Mn Si or Ni Cr V W Mo Co

Water-hardening tool steels


W1b 0.60-1.40a ... ... ...
146 Tool Design

W2b 0.60-1.40a ... ... ... 0.25


a
W3 0.60-1.40 ... ... 0.50
W4 0.60-1.40a ... ... 0.25
W5 0.60-1.40a ... ... 0.50
W6 0.60-1.40a ... ... 0.25 0.25
a
W7 0.60-1.40 ... ... 0.50 0.20
Shock-resisting tool steels
S1b 0.50 ... ... 1.50 ... 2.50
S2 0.50 ... 1.00 Si ... ... ... 0.50
S3 0.50 ... ... 0.75 ... 1.00
S4 0.50 0.80 2.00 Si
S5b 0.50 0.80 2.00 Si ... ... ... 0.40
Oil-hardening cold-work tool steels
O1b 0.90 1.00 ... 0.50 .... 0.50
O2 0.90 1.60

06 1.45 ... 100 Si ... .... 0.25

07 1.20 ... ... 0.75 ... 1.75 0.25 opt

Air-hardening medium-alloy cold-work tool steels


b, c
A2 1.00 ... ... 5.00 ... ... 1.00

A4 1.00 2.00 ... 1.00 ... ... 1.00


Contd...
Designation C Mn Si or Ni Cr V W Mo Co

A5 1.00 3.00 ... 1.00 ... ... 1.00

A6 0.70 2.00 ... 1.00 .... ... 1.00

A7 2.25 ... ... 5.25 4.50 ... 1.00

High-carbon high chromium cold-work steels

D1 1.00 ... ... 12.00 ... ... 1.00

D2b, c 1.50 ... ... 12.00 ... ... 1.00


b, c
D3 2.25 ... ... 12.00
b
D4 2.25 ... ... 12.00 ... ... 1.00

D5(b) 1.50 ... ... 12.00 ... ... 1.00 3.00

D6 2.25 ... 1.00 Si 12.00 ... 1.00


c
D7 2.35 ... ... 12.00 4.00 ... 1.00

Chromium hot-work tool steels

H11 0.35 ... ... 5.00 0.40 ... 1.50

H12b 0.35 ... ... 5.00 0.40 ... 1.50


b, c
H13 0.35 ... ... 5.00 1.00 ... 1.50
H14 0.40 ... ... 5.00 ... 5.00
H15 0.40 ... ... 5.00 ... ... 5.00
H16 0.55 ... ... 7.00 ... 7.00
Tungsten hot-work tool steels
Tooling Materials and Heat Treatment 147

H20 0.35 ... ... 2.00 ... 9.00


Contd...
Designation C Mn Si or Ni Cr V W Mo Co
b
H21 0.35 ... ... 3.50 ... 9.50
H22 0.35 ... ... 2.00 ... 11.00
H23 0.30 ... ... 12.00 ... 12.00
148 Tool Design

H24 0.45 ... ... 3.00 ... 15.00


H25 0.25 ... ... 4.00 ... 15.00
H26 0.50 ... ... 4.00 1.00 18.00
Molybdenum hot-work tool steels
H41 0.65 ... ... 4.00 1.00 1.50 8.00
H42 0.60 ... ... 4.00 2.00 6.00 5.00
H43 0.55 ... ... 4.00 2.00 ... 8.00
Tungsten high-speed tool steels
T1b, c 0.70 ... ... 4.00 1.00 18.00
T2c 0.85 ... ... 4.00 2.00 18.00
T3c 1.05 ... ... 4.00 3.00 18.00
T4 0.75 ... ... 4.00 1.00 18.00 ... 5.00
T5 0.80 ... ... 4.00 2.00 18.00 ... 8.00
T7 0.75 ... ... 4.00 2.00 14.00
T8 0.80 ... ... 4.00 2.00 14.00 ... 5.00
T15c 1.50 ... ... 4.00 5.00 12.00 ... 5.00
Molybdenum high-speed tool steels
M1b, c 0.80 ... ... 4.00 1.00 1.50 8.50
M2b, c 0.85 ... ... 4.00 2.00 6.25 5.00
M3b, c, d 1.00 ... ... 4.00 2.40 6.00 5.00
M4 1.30 ... ... 4.00 4.00 5.50 4.50
Contd...
Designation C Mn Si or Ni Cr V W Mo Co
M6 0.80 ... ... 4.00 1.50 4.00 5.00 12.00
M7 1.00 ... ... 4.00 2.00 1.75 8.75
M10b, c 0.85 ... .... 4.00 2.00 ... 8.00
M15 1.50 ... ... 4.00 5.00 6.50 3.50 5.00
M30 0.80 ... ... 4.00 1.25 2.00 8.00 5.00
M33 0.90 ... ... 3.75 1.15 1.75 9.50 8.25
M34 0.90 ... ... 4.00 2.00 2.00 8.00 8.00
M35 0.80 ... ... 4.00 2.00 6.00 5.00 5.00
M36 0.80 ... ... 4.00 2.00 6.00 5.00 8.00
Low-alloy special-purpose tool steels
L1 1.00 ... ... 1.25 ...
L2 0.50-1.10(a) ... ... 1.00 0.20
L3 1.00 ... ... 1.50 0.20
L4 1.00 0.60 ... 1.50 0.20
L5 1.00 1.00 ... 1.00 ... ... 0.25
L6 0.70 ... 1.50 Ni 0.75 ... ... 0.25 opt
L7 1.00 0.35 ... 1.40 ... ... 0.40
Carbon-tungsten tool steels
F1 1.00 ... ... ... ... 1.25
F2 1.25 ... ... ... ... 3.50
F3 1.25 ... ... 0.75 ... 3.50
Low-carbon mould steels
P1 0.10 max ...
P2 0.07 max ... 0.50 Ni 1.25 .... .... 0.20
Tooling Materials and Heat Treatment 149
Contd...
Designation C Mn Si or Ni Cr V W Mo Co
P3 0.10 max .... 1.25 Ni 0.60
P4 0.07 max .... .... 5.00
P5 0.10 max .... .... 2.25
150 Tool Design

P6 0.10 .... 3.50 Ni 1.50 0.20


P20 0.30 ... ... 0.75 .... .... 0.25
PPT 0.20 1.20 Al 4.00 Ni
Other alloy tool steelse
6G 0.55 0.80 0.25 Si 1.00 0.10 .... 0.45
6F2 0.55 0.75 0.25 Si, 1.00 0.10 opt ... 0.30
1.00 Ni
6F3 0.55 0.60 0.85 Si, 1.00 0.10 opt ... 0.75
1.80 Ni
6F4 0.20 0.70 0.25 Si, ... ... ... 3.35
3.00 Ni
6F5 0.55 1.00 1.00 Si, 0.50 0.10 ... 0.20
6F6 0.50 ... 2.70 Ni 1.50 ... ... 0.20
6F7 0.40 0.35 4.25 Ni 1.50 ... ... 0.75
6H1 0.55 ... ... 4.00 0.85 ... 0.45
6H2 0.55 0.40 1.10 Si 5.00 1.00 ... 1.50
a
Various carbon contents are available in 0.10% ranges.
b
Stocked in almost every warehousing district and made by the majority of tool-steel producers.
c
Available as free-cutting grade.
d
Available with vanadium contents of 2.40 or 3.00%.
e
The designations w of these steels are similar to those used in the 1948 w “Metals Handbook,” expect they were previously written with Roman numerals (V1 F2, etc). Neither
AISI nor SAE has assigned type numbers to these steels.
SOURCE: ASM, “Metals Handbook”, 8th ed., vol. 1, 1961, by permission.
Tooling Materials and Heat Treatment 151

heat softening. They are easy to machine compared to other steels and require rela-
tively simple heat-treating methods. They are suitable for light or medium cold impact
operations such as coining, cold heading, punching, knurling, embossing, and for
wood and metal hand cutting tools. They are probably the best all-round tool steels
under low-temperature working conditions.
The water-hardening tool steel should not be used when the tool has drastic dimen-
sional changes, sharp corners, or holes near the edge of the tool. Tools of this nature
have a tendency to crack during heat treatment because of quenching stresses
caused by uneven cooling and stress concentration points. They should not be used
when high working temperatures will be encountered, as the tool will soften. They
should not be used where distortion during heat treatment would present a problem.
Group W steels have a tendency to warp during quenching.
When the tool is relatively complicated and there is a question whether the tool will
crack during heat treatment, the W group of tool steels should not be used. The added
expense of oil-or air-hardening grades may pay in this situation.

Shock-resisting tool steels: group S These steels are used for stock operations at
normal temperatures and where maximum abrasion resistance is not required. They
contain less carbon and have higher toughness, and hardness is usually below RC60.
They are oil-and water-hardening steels.
The chief alloying elements are silicon, chromium, tungsten, and molybdenum. These
alloys increase hardenability and provide heat and wear resistance. The high-tung-
sten types are characterised by higher heat resistance.

Oil-hardening cold-work steels: group O Group O tool steels are one of the more
important groups in that they are able to overcome some of the difficulties encoun-
tered in the use of water-hardening steels (group W). They lend themselves to safer
hardening and have less dimensional change during heat treatment. They are rel-
atively inexpensive, readily available, have good machinability, good resistance to
decarburisation, and have a high enough carbon content to provide good wear resis-
tance. The depth of hardening is greater than that of water-hard steels, and as a result
they are usually less tough. They do not have high red hardness and therefore must
be used for tools that will operate near room temperature.
These steels have a wide range of applications. Specific examples are blanking,
bending, trimming, coining, shearing and shaping dies, thread-rolling dies, broaches,
knurling tools, gauges, and other application beyond the scope of carbon grades.

Air-hardening cold-work steels: group A Air-hardening tool steels have the advan-
tages of oil-hardening tool steels but to a greater extent. Manganese, chromium,
molybdenum, and vanadium are the chief alloying elements, and their main func-
tion is to promote air-hardening characteristics which result in excellent dimensional
stability. The high carbon content provides good wear resistance, and the high-man-
ganese grades may be hardened at lower temperatures, thus reducing scalling and
further reducing dimensional change. In general, machinability is not as high as in the
water- and oil-hardening grades.
152 Tool Design

The air-hardening grades are applicable to intricate tool shapes such as thread-rolling
dies, long slender broaches, dies with projections, and other applications where resis-
tance to distortion and abrasion is of prime importance.

High-carbon-high-chromium cold-work steels: group D Group D tool steels com-


bine high wear resistance with deep hardening properties. These characteristics are
caused by the high content of chromium and carbon. They have extremely low dimen-
sional change during hardening and have a medium resistance to heat softening.
They are susceptible to edge brittleness, which makes them unsuitable for edge-
cutting tools.
Specific applications for group D steels are wire-drawing dies, master gauges, intri-
cate blanking and piercing dies, and other applications where dimensional stability
and long-wearing properties are important.

Hot-work tool steels: group H The tool steels discussed to this point have from
low to medium resistance to heat softening. Group H tool steels have been alloyed
to withstand high working temperatures for such applications as hot-forging dies, hot-
extrusion dies, hot shears, die-casting dies, and plastic-moulding dies. The main alloy-
ing elements for this group of steels are chromium, molybdenum, and tungsten. The
hot-work tool steels are divided into three groups, depending upon the major alloying
element.

Chromium hot-work tool steels (group H11 to H16) These contain from 5 to 7
per cent chromium and smaller amounts of vanadium, tungsten, and molybdenum.
In addition to good red-hardening properties, they are extremely deep hardening and
have good dimensional stability during hardening.

Tungsten hot-work steels (group H20 to H26) Containing from 9 to18 per cent
tungsten and 2 to 12 per cent chromium these alloys increase in resistance to high-
temperature softening and washing compared to grades H11 to H16. They are, how-
ever, more brittle at working hardness and generally cannot be successfully water
cooled during service.

Molybdenum hot-work steels (group H41 to H43) These contains 5 to 9.5 per cent
molybdenum, 4 per cent chromium, 1.5 to 6.5 per cent tungsten, and smaller amounts
of vanadium. They have properties almost identical to those of the H20 to H26 steels,
their main advantage being a lower initial cost.

High-speed tool steels The high-speed tool steels have a high degree of red
hardness and high abrasion resistance along with a comparable degree of shock
resistance. Their primary use is as a material for cutting tools, although they have
other applications, such as extrusion dies and blanking punches and dies. Their
major alloying elements are tungsten, molybdenum, chromium, and vanadium, and
in special grades, cobalt is added to give superiority in red hardness and abrasion
resistance. High-speed steels are more difficult to machine and grind because of the
high carbon and alloy content.
The high-speed steels have been divided into tow groups: (i) tungsten high-speed
steels, group T, and (ii) molybdenum high-speed steels, group M. There is little differ-
Tooling Materials and Heat Treatment 153

ence in the characteristics of the two groups. The main deference is in the initial cost,
group M being approximately 30 per cent less for equivalent grades.
The general-purpose grades of high-speed steel are Y1, M1, M2, and M10. When
the highest possible red hardness is required in a cutting-tool material, a cobalt high-
speed steel such as M6 may be used. However, the cobalt high-speed steels are
higher in cost and are difficult to machine, heat treat, and grind.

Special-purpose tool steels A number of tool steels that do not fall into the basic
tool-steel categories have been divided into the following groups:
Low-alloy special-purpose tool steels, group L These steels are similar to the water-
hardening steel (group W) with the addition of chromium and other elements for
greater wear resistance and hardenability. The L6 grade has additions of nickel for
increased toughness and hardenability. They are used where high wear resistance
and toughness are required, as in bearing, chuck parts, and indexing fingers.
Carbon-tungsten tool steels, group F These water-hardening steels have high wear
resistance because of the high carbon and tungsten content. They are used for high
wear, low temperature, and low-shock applications, as in wire drawing dies, paper-
cutting knives, forming tools, and burnishing tools.
Low-carbon mould steels, group P Group P steel are alloy carburising steels pro-
duced to tool-steel quality. They have a low hardness in the annealed state and are
generally carburised for greater wear resistance after being machined. Their major
application is moulds for injection or compression moulding of plastics.

3.4 CAST IRON


Gray cast iron is sometimes used as the main body of jigs and fixtures. It is some-
times easier to cast a shape than to build it up with several places of steel. Metal may
be placed to better advantage, and as a result the weight of the fixture or jig may be
reduced. The stability and compressive strength of gray cast iron, as well as its ease
of casting, make it suitable for this purpose.
Cast iron is also used in the construction of large forming and drawing dies and as a
material for die-set shoes.

3.5 MILD OR LOW-CARBON STEEL


Low-carbon steels are used extensively by the tool designer. Hot and cold-rolled flats,
cut to size and properly machined, are used as component parts of jigs and fixtures
where wear resistance and maximum strength are not a necessity. Standard struc-
tural shapes are used in construction of frameworks for large jigs and fixtures.
Cold-roller shapes are smoothly and accurately finished and are commonly used for
component part that require little or no machining. Hot-rolled shapes have an oxide
scale and are not finished as accurately as cold-rolled but are lower in cost and there-
fore may be used when extensive machining is necessary. Hot-rolled shapes are
also better suited when the tool is fabricated by welding, as cold-rolled steel has a
greater tendency to warp when heated or machined due to the extra stresses set up
during the cold-rolling operation at the steel kill, which are relieved during heating or
machining.
154 Tool Design

3.6 NONMETALLIC TOOLING MATERIALS


Rubber This material, in the form of a rubber pad, is used in various specialised
forming, drawing, blanking, and bulging operations. It is also used as stripper material
with conventional punches and dies.

Masonite This is a cellulose material that is sometimes used for the construction of
punches and dies in drawing operations on thin gauges of metal. Typical examples
would be form blocks in rubber forming and stretch dies. It can be used for blanking
and punching operation providing steel inserts are used to provide cutting surfaces.
Its advantages lie in its lightweight, ease of working, ease of repair, and the fact that
it does not scratch the finished part.

Densified wood various woods are laminated and impregnated with phenolic resin
and compressed to about 50 per cent of the original thickness of the wood layers.
Dandified woods are used in the construction of forming and drawing dies for soft
materials. They are also used for the die-board components of steel-rule dies.

Plastics The aircraft industry is responsible for the major developments of plastics
as tooling materials, because of short lead time and limited production runs of aircraft
parts.
The advantages of plastics as tooling materials are reduced labour costs, reduced
lead time, ease of tool-design changes, easy revision and repair, inexpensive fabri-
cating equipment, ease of duplicaton, and resistance to moisture, temperature, and
chemicals. The choice between plastics and other conventional tooling materials gen-
erally depends upon the severity of operation and the length of the production run.
The major plastics used in various types of tooling are phenolics, polyesters, ure-
thanes, and epoxies. They are suited for an unlimited number of applications, such
as stretch dies, hydraulic-press dies, draw dies, and duplicator pattern. They can be
laminated with other materials, surface-cast, mass-cast, or used as paste plastic tool-
ing. Chapter 11 is devoted to the subject of plastic tooling.

Oxide cutting tool materials As a tooling material, the application of oxides is gen-
erally limited to cutting tools. It is a relatively new cutting-tool material, having been
brought to public attention in 1955. Many improvements have been made since its
introduction, and for certain machining operations it has distinct advantage. It does,
however, have distinct limitations. As it is one of the newer cutting tools, it will con-
tinue to improve in the future.
At the present time, oxide cutting tools are basically composed of aluminium oxide,
Al2O3. It should be noted that this is the same material used in aluminium oxide grind-
ing wheels. Some are bonded together with metallic binders, while others are fused
together without the use of other elements. They were originally referred to as ceram-
ics, but more recently the term oxides is being used, as it is more descriptive of
the main ingredients. The main characteristics of oxide tools are high compressive
strength, high hot hardness, high abrasive resistance, low heat conductivity, resis-
tance to galling and welding. Compared to cemented carbides, oxides are generally
slightly harder, but they are considerably weaker when exposed to bending and ten-
sion loads. As a result, tools must be designed so that the oxide is loaded in com-
pression along with insert shapes that are geometrically blocky. Oxides can be run
Tooling Materials and Heat Treatment 155

at much higher speeds than carbides because of their oxidation resistance at high
temperatures. They have been run at 5500 m/min under laboratory conditions, but
speeds of 180 to 500 m/min are typical. When cutting speeds are concerned, oxide
cutting tools begin where cemented carbides leave off.
The low strength of oxide cutting tools is their chief limitation. They are not suitable
for shock loads of interrupted cuts. For this reason, they have not yet been applied
successfully for milling operations. It is not possible to braze inserts successfully to
steel shanks because of their nonwetable properties. This, of course, is an advantage
when cutting metal, as there is a reduced tendency for chips to weld to the tool. Oxide
inserts are held mechanically or by resin bonding. Resin bonding is possible because
of the low heat conductivity of the oxide insert.
Another factor to be considered when using oxides as cutting tools is that machine
tools must have greater power, speed, and must be in excellent working condition.
There can be no slip in the bearings, slides, and driving and feeding mechanisms
of the machine. The tool-and work-holding mechanisms must be rigid and free from
vibration.
The present practical applications of oxide cutting tools are in the machining of cast
iron, carbon, and low-alloy steels; finishing of hard steels (RC60 to 65) at high speeds,
and other highly abrasive nonmetallics.

Diamond Although limited in their application, diamonds as a tooling material


should be mentioned. Diamond powder is used for grinding and polishing, while
whole diamonds are used as turning tools, grinding-wheel dressers, and inserts for
wire-drawing dies.
Industrial diamonds are either natural or man-made. They are considered to be the
hardest known substance. Other characteristics are high abrasion resistance, chemi-
cal inertness, high strength, high heat conductivity, melting point, high modulus of
elasticity, and low compressibility.
Diamonds are used in precision production turning of plastics, precious metals, light
metals, and difficult-to-machine materials. They are used in place of other cutting tool
materials because their hardness and lasting sharp cutting edges reduce deflection
forces and cut through the grain structure of the material. This results in less surface
smear, an exellent surface finish, and extremely close tolerance work. They are gen-
erally used only as finishing tools with light cuts, fine feeds, and high cutting speeds.
Depths of cut may be from 25 to 13 mm, with tolerance of 1.3 mm and closer being
held. The machining speed is usually as high as the machine will operate without
vibration.

3.7 NONFERROUS TOOLING MATERIALS


Sintered carbides The sintered, or cemented, carbides are basically hard carbides
of tungsten, titanium, and tantalum held together with a cobalt binder. Other materials
may be involved, such as chromium and nickel. They are products of powder metal-
lurgy, which fundamentally consist of mixing powder carbides with powder cobalt in
a ball mill, followed by pressing or compacting the mixed powder into the desired
shape. The resulting blank is then presintered to increase the hardness so that it can
156 Tool Design

easily be handled and formed. It is then formed by various machining methods and
final-sintered to complete the process.
Sintered carbides are characterised by high hardness (RA85 to 93), high compres-
sive strength, and high red hardness. They are used for metal cutting more than any
other operation. Cutting speeds and feeds are generally higher than those used with
high-speed steel and cast-alloy cutting –tool materials. For example, steels may be
successfully machined from 30 to 300 m/min.
They are broadly classified in two groups: (i) the strains tungsten carbide grades,
containing tungsten carbide and cobalt along with small amounts of titanium and tan-
talum, and (ii) the steel-cutting grades containing larger amounts of titanium and tan-
talum carbides. The straight tungsten carbide cutting grades are used for machining
cast iron, nonferrous alloys, and nonmetallic materials. The steel-cutting grades are
used for machining steels and steel alloys. The heavy additions of titanium tantalum
prevent the steel chip from causing crate ring (washing out of carbide particles from
behind the cutting edge).
In addition to metal cutting, sintered carbides are used for wire-drawing dies, long-
run blanking dies, mandrels, gauges, wear pads, boring bars, and other applications
where strength, rigidity, and resistance to heat and wear are necessary.

Cast nonferrous alloys Sometimes referred to as cast alloys, these tool materials
are chiefly composed of cobalt, chromium, tungsten, and carbon. They are usually
cast to the desired shape and size, and thus are named cast alloys. High red hard-
ness and abrasion resistance are characteristics of these alloys. They do not have to
be heat treated, as hardness in the as-cast condition is RC60 to 65. The hard grades
are primarily used as cutting tools and are able to withstand cutting speeds twice as
high as employed with high-speed cutting tools. They are weaker and more brittle
than high speed tools but are able to withstand more shock than carbide cutting tools.
The result is that they serve as a cutting material with characteristics intermediate
between high-speed steel and sintered carbides.
Cast alloys are available commercially in a variety of sizes and shapes. They may
be ground with the same wheels used for grinding high-speed steels. Since they are
more costly than high-speed steel, the tools are not made entirely of the alloy materi-
als except in the smaller sizes. Cast-alloy inserts may be brazed, welded, or mechani-
cally attached to mild-steel shanks.
Other applications for cast alloys are burnishing rollers, wear strips, antifriction bear-
ings exposed to high heat, and hard facing of dies and gauges.

Zinc-base alloys These alloys may be used as materials for punches and dies for
short-run production of aluminium sheet up to 1.5 mm thick. The main advantage of
this application is the ease and speed of fabrication. They may be used in the form
of tooling plate or may be cast. Cast tools form well to the desired shape and require
little machining. When used as punch and die materials, zinc-base alloys tend to flow
toward the cut, and in this respect tend to be self-sharpening.

Bismuth alloys The bismuth, or low-melting, alloys are alloys of bismuth, lead, tin,
cadmium, and other matals. They are referred to as low-melting alloys because the
melting points may be below the boiling point of water. High-bismuth alloys expand
when they solidify, which makes them excellent material for duplicating mould detail.
Tooling Materials and Heat Treatment 157

Bismuth alloys are available commercially in a number of compositions, depending


upon the characteristics desired. Specific applications are moulds for duplicating
plaster or plastic patterns, anchoring punches and dies for blanking, chucks for hold-
ing special or engraving and duplicating.

Magnesium The important characteristic of magnesium compared to other materi-


als is its lightweight. For this reason, it is used as structural material for large assem-
bly fixtures in the aircraft industry. The main advantage is that fixtures are easier to
handle. Another advantage is that it does not deflect under its own weight, as many
heavier materials do. It is also easily machined and welded. Its thermal expansion is
almost the same as aluminium, which make it particularly well suited for the aircraft
industry.
Magnesium is also used as tooling plate and facing material for forming blocks.

3.8 HEAT TREATING


The tool designer is usually not responsible for the heat treating of tool steels; how-
ever, he should have a good knowledge of heat-treating capabilities of his company,
as they must be taken into consideration as he designs various types of tools. The only
way to fully understand his company’s heat-trating capabilities is to have a through
understanding of heat-treating processes.
Before beginning the discussion of actual heat-treating practises, it would be well to
define common terms that are associated with the heat treatment of tool steels.

Heat treatment Heat treatment, or heat treating, is a process whereby the physical
properties of a metal are changed by subjecting it to a combination of heating and
cooling. The purpose may be to harden, soften, toughen, stress-relieve, increase
machinability, increase strength, or a combination of these. The degree and rate of
heating and cooling will depend upon the properties desired.

Normalising The purpose of normalising is to put a ferrous material back into a nor-
mal structure after forging, casting, or improper heat treatment. The process results in
grain refinement, homogeneity of the structure, and, in some case, increased machin-
ability. These properties put a steel into a condition that enable it to respond correctly
to further heat treatments.
Normalising is accomplished by heating the steel to approximately 40oC above the
usual hardening temperature and allowing it to cool in still air.

Annealing The purpose of annealing is generally to soften, although the term is


sometimes used when the purpose is to stress relieve. A tool steel may be too hard
to machine as the result of the hammering, rolling, or improper cooling at the mill;
or a tool may have been previously hardened and require annealing for additional
machining.
Annealing for the purpose of softening is accomplished by heating the steel to a tem-
perature slightly above its hardening temperature and cooling slowly in the furnace
or should be carried out to low temperatures. Exact annealing temperatures may be
obtained from the manufacturer of the particular type of tool steel.
158 Tool Design

Spheroidising Carbon steel with a high percentage of carbon (one per cent and
over) may be difficult to machine even though they have been annealed. This is due
to an excess of hard plates of iron carbide. Iron carbide is quite hard, and the cutting
tool is required either to cut through the iron carbide plates or push between them in
order to remove metal. A heat-treating process which will improve the machinability
of a steel in the length of the carbides and process which will improve the machin-
ability of a steel in the length of the carbides and produces a spheroidal or globular
form. This structure is easier to machine, as the tool can get between the carbide
more easily.
Spheroidising is accomplished by prolonged holding at a point just below the lower
critical temperature or by heating and cooling alternatively between temperatures that
are just above and below the lower critical temperature.

Critical Temperature The critical temperature, or critical point, is the temperature


at which various transformations occur in steel occur in steel as it passes through its
critical range. In order for the steel to harden, it muss be heated above this tempera-
ture. The critical temperature varies with the analysis of the steel.

Critical cooling rate The critical cooling rate is the quenching speed needed to
harden a steel.

Hardenability The hardenability of steel is its ability to harden when quenched from
its hardening temperature, as measured by its surface hardness and by the depth of
hardening below the surface. As explained earlier, water-hardening tool steels have
a hard shell and tough core after hardening. This is because they must be quenched
very rapidly in order to reach maximum hardness. The hard shell and tough core
result because the inside cools slower than the outside. Oil-hard steels harden when
cooled more slowly, thus the hardness penetration or hardenability, is greater.

Stress relieving Stress relieving consists of heating the steel to a suitable tempera-
ture, holding it long enough to reduce residual stresses, and following by slow cooling.
For example, if a large amount of metal is machined from a tool, a certain amount of
stress is set up as the result of the machining operation. When heat treated, the tool
may warp because of the strains produced when the stresses are relieved. This can
be prevented by rough machining the tool, giving the tool a stress-relieving heat treat-
ment, and finally machining the tool to finished size.
The stress-relieving temperature is slightly below the critical point of steel. Furnace
cooling is preferred; however, cooling in an insulating material such as lime of ashes
is acceptable.

Stabilising Heat-treated steels have a tendency to change dimensions slightly over


a long period of time. This dimensional change may be very small, but in the case of
very accurate gauges, this change may be detrimental. To combat this slight dimen-
sional change, tools may be artificially aged, or stabilised. This is accomplished by
hardening and tempering the tool and before final lapping the tool is heated in boiling
water, followed by cooling naturally back to room temperature. The tool is then “fro-
zen” in dry ice or suitable deep-freeze equipment and again allowed to return to room
temperature. This cycle of heating and cooling is repeated four or five times. The tool
is then lapped to finish size. Stabilising as described can have the same effect as 10
to 20 years of natural aging.
Tooling Materials and Heat Treatment 159

Hardening Tools are hardened to give them strength and wear resistance. Hardening
is accomplished by heating the tool to above its hardening temperature (supplied by
the manufacturer of the tool steel) and quenching it in the proper medium. Hardening
will be discussed in detail later in this chapter.

Pack hardening Pack hardening is a process whereby the tool to be hardened is


packed in a material to protect it from the atmosphere and prevent scaling and decar-
burisation. This process is often used when a hardening furnace with a controlled
atmosphere is not available. The tools are generally packed in a container with clean
cast-iron chips, and the entire container is placed in the furnace. The pack is then
heated to the hardening temperature. Sometimes a thermocouple is placed in the
pack near the part to determine when the part reaches the hardening temperature.
Cast-iron chips are used because they add little or no carbon to the surface of the
part.

Quenching Quenching is cooling a heated piece of metal. The quenching medium


refers to the substance in which the metal is cooled. Common quenching media are
brine, water, oil, caustic soda in water, molten salts, and still air. Quenching will be
given thorough consideration later in this chapter.

Tempering Tempering, sometimes known as drawing, is a process whereby a cer-


tain degree of hardness is sacrificed to relieve strains and increase toughness. This
is accomplished by reheating the steel after hardening to a temperature much lower
than the hardening heat. The length of time required to draw or temper a tool depends
upon its size and shape.
Tempering may be done in baths of oil, salts, or lead, whose temperatures are pyrom-
eter-controlled. A box-type furnace with controlled temperature may also be used.
The advantage of liquid baths over box-type furnaces is that they heat faster and the
temperature is uniform throughout the bath.
It should be noted that although a tempered tool will lose some of its hardness, the
purpose of tempering is not to decrease hardness. The purpose of tempering is to
decrease internal strains and increase toughness. For instance, in the heat treat-
ment of metal-cutting tools, it would be desirable to have the tool as hard as possible.
However, if the tool was not tempered, the internal strains set-up by hardening could
cause the tool to crack at the slightest shock load. Therefore, they are tempered at a
relatively low heat only relieve strains. Hardened tools that have not been tempered
have been known to crack when left on the bench overnight.
An example of tempering for toughness would be in the heat treatment of pneumatic
tools. Here the tool is exposed to heavy shock loads, and maximum toughness is
desired. Tempering would be performed at a higher temperature than in the heat treat-
ment of the previous cutting tool to give maximum toughness.

Double tempering Higher alloyed tool steels (high-speed and hot-work steels) that
are hardened from high temperatures are double tempered. This is to achieve the
maximum effect of the tempering treatment. These steels retain austenite tin the
quench, and fresh toughen the newly formed martensite formed in the first temper.1
The second temper is usually 25 to 50o lower than the first temper.

1
For a definition of austerite and marterisite, see Sec. 3.9.
160 Tool Design

Decarburisation Decarburisation is the loss of carbon on the surface of ferrous


alloys. This is the result of oxygen reacting with the surface of the metal when it is
hot. Tool steels have a decarburised surface when they are received from suppliers.
This surface is sometimes referred to as bark and must be removed before the steel
is usable as a tooling material. If it is not removed, the tool surface would be soft after
heat treament because of the lack of carbon. Failure to remove bark may also result
in cracking during heat treatment. The process of removing the decarburised surface

Table 3.5 Amount of material to be removed to eliminate decarburized surface of tool


steel*
Dimension of bar, mm Minimum amount to be removed per side, mm

Up to 12.5 0.40
Over 12.5 to 32 0.80
Over 32 to 76 1.6
Over 76 to 127 3.2
Over 127 4.8
*For rectangular sections use the machining allowances shown when the width is less than four time the thick-
ness. when the width is four times the thickness or wider, use double the machining allowances for the thickness
dimension.
SOURCE: Carpenter Steel Company, “Tool Steel Simplified,” 2d ed., 1960, by permission.

by use of a machine tool is called barking. Table 3.5 shows the amout of material to
be removed by matching in order to remove the decarburised surface.
Many tools-steel companies offer prefinished decarburised-free (DCF) flats that do not
require barking to remove the decarburised surface. DCF flats and squares require
little or no preparation before layout. Substantial savings may accrue from the use of
these materials.
Decarburisation is also of concern when hardening finished parts in furnaces that
have no means of controlling the furnace atmosphere. Parts will have a “soft” surface
because of the loss of carbon. The control of decarburisation will be discussed later
in this chapter.

Microstructure Microstructure refers to the structure of polished and etched met-


als as revealed by a microscope. The study of microstructure is called microscopy
or metallography. The microstructure of tool steels depends primarily upon the heat
treatment to which it has been subjected. A trained metallurgist is able to observe
the microstructure of a tool steel and tell whether it has responded properly to heat
treatment.
The various structures of carbon steels are called cementite, ferrite, pearlite, austen-
ite, and martensite. These terms will be explained in Section 3.9. The tool designer
should be familiar with and in trade journals related to his profession.

Preheating Preheating sometimes precedes the regular heat-treating procedure,


especially in the case of higher alloyed tool steels such as high-speed steel. The
purpose is to reduce the thermal shock resulting when a cold tool is placed directly
into a very hot furnace and to arrange the carbides in a more favourable condition for
dissolving at the higher temperature. Reducing thermal shock helps prevent warp-
ing and cracking, Preheating also serves to relieve machining stresses. Preheating
Tooling Materials and Heat Treatment 161

temperatures vary from 650 to 870oC, depending upon the type of tool steel being
treated.

3.9 APPEARANCE OF CARBON IN STEEL


To better understand heat treatment and what actually causes metal to react to heat
treatment, one needs to know the various constituents of steel as they appear at vari-
ous temperatures. Carbon steel will be used be used as an example, and although the
various tool-steel alloys may react somewhat differently, carbon steel will serve as a
hosts for general understanding.
At room temperatures, carbon in steel occurs as iron carbon (a compound of iron and
carbon, Fe3C which contains 6.67 per cent carbon by weight). It is hard, brittle, has
low tensile strength, and high compressive strength. It is this substance which causes
difficulty in machining high-carbon steels. Low-carbon steels contain iron carbide but
not in such quantity as high-carbon steels. Metallurgists call the iron carbide constitu-
ent cementite.
Along with cementite, there exists in steel a soft substance composed of nearly pure
iron. Metallargists call this substance ferrite. It is the principal constituent of all steels
and the predominant one in low-carbon steels. When the carbon content of steel is
low, very little cementite is formed and the structure will appear very much like pure
iron. As the carbon content increases, more cementite is formed in narrow layers
and areas of cementite and ferrite are called pearlite. Under a microscope, it has the
appearance of mother-of-pearl, from which its name is derived.
As the amount of carbon is increased, the pearlite areas increase and the amount of
ferrite decreases unttil 0.85 per cent carbon is reached. At this point the structure is
all pearlite. When the carbon is higher than 0.65 per cent, and envelope of cementite
surrounds the pearlite and appears a white networks after etching (see Fig. 3.1 for
micrographs showing these constituents).

(a) Annealed 1020 steel (b) Annealed 1040 steel

Fig. 3.1 Micrographs showing the common constituents of carbon steel: (a) annealed
1020 steel (b) annealed 1040 steel (The Carpenter Steel Company)
162 Tool Design

As steel is heated, no significant change takes place in the structure until a tempera-
ture of about 720°C is reached. At this temperature the cementite begins to go into
solid solution with the ferrite. The formation of this solution is similar to that of a solu-
tion of salt in water, except that the cementite and ferrite are both solid. As further
heating takes place, the solid solution grows until all the cementite and ferrite disap-
pear. The iron carbide, instead of being found in pearlite grains, is distributed uni-
formly throughout the entire solution, and any small area analyses the same as any
other. When viewed through a metallurgical microscope, the new structure (the solid
solution) has a bleak appearance similar to that of a pure metal. This new structure is
known to metallurgists as austenite.
The range of temperature within within which the austenite is formed is referred to
as the critical range or transformation range. The limiting temperatures of the ranges
depend upon the amount of carbon in this example or in the composition of the alloy
in tool steel. Lower carbon content, the wider the transformation range will be. Upon
increasing the carbon content, the transformation range will become narrower until
the carbon content reaches 0.83 per cent. At this point the limiting temperatures
merge and become one, and the change to a solid solution will occur at one time. This
temperature is generally given as 723°C.
Further heating above the transformation range causes no other change in steel
except to enlarge the size of the grain. Higher the temperature and longer the steel is
held at this temperature, coarser the grain will become.
If the steel is allowed to cool slowly through the transformation range, the reverse
will take place but at a temperature of about 5° lower. The structure will be the same
as in the beginning, with the exception of the grain size. The maximum temperature
reached above the transformation range determines the final effective grain size. This
point should be remembered, as a tool may be spoiled by heating it too high above
the transformation range during heat treatment. This would result in a very coarse
grain structure that would be detrimental to the tool.
Steel in the solid state is always made up mostly of crystalline grains, but at different
temperatures the internal structure of the crystals is not always the same. This phe-
nomenon is known as allotropy, which means that certain metals may exist in more
than one crystalline structure. For our purposes, it is only necessary to understand
that the crystals of steel exist in one structure above the transformation range and in
an entirely different structure below the transformation range. Transformation in iron
and steel depend upon transitions from one allotropic form to another. Metallurgists
refer to the structure above the transformation range as the gamma structure, while
the structure below the transformation range is know as the alpha structure.
When this change of structure occurs, i.e., from alpha to gamma, certain physical
changes may be observed. If a thermocouple (a high-temperature measuring device)
is placed within a block of steel heated, one will observe a check in the rise of tem-
perature as the steel passes through the transformation range. This would indicate
that a physical change is taking place within the steel, as energy in the form of heat is
being absorbed. Upon slow cooling back through the transformation range, there will
be a rise in temperature, indicating the reverse has taken place.
Another change which can be observed as the steel block goes through the trans-
formation range is the loss of magnetism. A magnet that is attached to the block
Tooling Materials and Heat Treatment 163

at room temperature drops off when the steel is heated through the transformation
range. Noting the loss of magnetism by use of a permanent magnet has been used
as a method to determine when steel has reached its hardening temperature, but it
is not reliable, as the hardener cannot be sure that the whole body of steel has been
completely transformed.
The third physical change is that of volume. It is commonly known that when a steel
block is heated it will expand evenly in all directions. This change occurs uniformly
with the increase in temperature until the steel reaches the transformation period. As
the steel goes through the transformation period, there is a decrease in volume. This
is due to the rearrangement of the atoms of the individual crystals. This change in
volume has a great effect on warping and cracking during heat treatment.
The effect of slow heating and slow cooling of steel has been discussed. To under-
stand why steel reacts to heat treatment, it will be necessary to consider rapid cooling
of steel.
As previously described, steel in the austenitic condition is a uniform solid solution
existing in crystalline form. If it is suddenly cooled through the transformation range, it
will retain its austentic structure at temperatures much below those at which it would
be retained at a slower cooling rate. This is because insufficient time is allowed for the
iron carbide to diffuse out of the austenite naturally, and the structure cannot revert
back into the original alpha structure that existed below the transformation range.
After cooling to room temperature, the result would be that the iron carbide would
come out of solution in extremely small particles and form a needlelike structure dis-
tributed evenly throughout the alpha iron. This new structure is known as martensite.
It is very hard, brittle, strong and is the constituent of fully hardened steel. The degree
of hardness, strength and brittleness depend upon the carbon content only. Hardness
is due to the extremely fine grain of the alpha structure along with finely divided hard
iron carbide particles dispersed in it.
This phenomenon of hardening during quenching takes place in two stages. The first
occurs when the steel cools rapidly from the austenising temperature to about 205 to
260°C. If the steel is not cooled rapidly enough through this range, softer constituents
will form and the steel will not reach maximum hardness. If the steel has been cooled
fast to about 205 to 260°C, softer constituents will not form and the steel will have to
harden, whether the tool cools further in the cooling bath or is pulled from the bath
and allowed to cool in still air. At this time the tool is relatively soft and can be bent.
Experienced hardeners take advantage of this fact and quickly pull tools from the
quench and straighten them before they begin the final stage of hardening. The trick
is to know when to pull them from the bath.
The second stage or actual hardening takes place at the end of the quenching cycle
and occurs at quite low temperature. These low temperatures can be obtained by
leaving the tool in the quenching bath during the entire quenching cycle or by pulling
the tool from the quenching bath at about at about 260°C and allowing it to cool in air.
This is providing that the tool has cooled fast enough to the 260°C point.
It is therefore necessary to make sure that the tool cools to a point low enough to com-
plete its entire hardening cycle before subjecting it to tempering treatments. Generally,
a tool has completed its hardening cycle when its temperature has dropped below the
boiling point of water.
164 Tool Design

Matensite, in the as-quenched condition,


is generally too brittle for most applica-
tions. This is due to high residual stresses
set up during quenching. It is therefore
necessary to temper the steel, which con-
sists of reheating it to some temperature
below the transformation range.
Reheating below 205°C does not seem
to modify the martensitic structure notice-
ably, but it does provide a great deal of
stress relief. Reheating above 205°C
allows gradual transformation of the mar- Fig. 3.2 Martensite, normal treated struc-
tensitic needlelike structure to a more ture, W1 tool steel (The Carpenter
granular form. Higher the temperature, Steel Company)
larger the granular particles become.
This type of structure is generally referred to as tempered martensite. A micrograph
of martensite is shown in Fig. 3.2.
So far in this discussion of heat-treating operations, the subject of specific tempera-
tures for the various operations has been avoided. This is because each steel has its
own heat-treating temperatures and procedures, which have been determined by its
manufacturer. The manufacturer’s recommendations usually are general in order to
try to cover as many application as possible. This is not to say that the manufacturer’s
recommendations cover all applications.
A person with a great deal of heat-treating experience may have sufficient knowledge
to enable him to depart from the manufacturer’s recommendations. His company
may be large enough to experiment with various types of steels and determine the
best methods and heat-treating temperatures. Smaller companies, however, may not
have the time or capital to experiment and therefore must rely on the manufacturer’s
recommendations.

3.10 FACTORS AFFECTING HEAT TREATING


Each tool-steel manufacturer has spent a great deal of time and money in the develop-
ment of his particular tool steel. Each company usually refers to his tool-steel products
by brand name, and although each brand name falls into various AISI classifications,
this does not mean every tool steel listed under an AISI classification reacts the same
to the same kind of heat treatment. This is another reason for relying on the tool-steel
manufacturer’s recommendations. If the hardener follows the recommendations to the
letter and heat treating fails to produce the required properties, he should immediately
call in the manufacturer’s representative, rather than experiment with different heat-
treatment temperatures and procedures. The manufacturer is better able to cope with
heat-treating problems, as this is his business. If he cannot solve the problem on the
basis of his past experience, he certainly will want to add the new challenge to his
background of experience.
The tool-steel business is highly competitive, and the manufacturer will want his prod-
uct to satisfy the user in every respect. He cannot afford to put a poor product on the
market, as there are too many tool-steel companies that produce high-quality prod-
ucts. The manufacturer can be trusted, in fact, he must be trusted.
Tooling Materials and Heat Treatment 165

Tool-steel companies publish heat-treating guides for their products obtainable free
of change Instructions for heat treating are given for each different type of steel. After
studying the instructions for various steels, it becomes apparent that heat treating
boils down to questions of temperature, time, atmosphere, and cooling rate. The fol-
lowing discussion will be concerned with these questions.

Temperature If the hardening temperature is too low, the tool-steel will not harden
properly or may harden only on the outer portions. On the other hand, if it is too
high, excessive grain growth, scaling, decarburisation, and warping may be the result.
Coarse grain causes brightness, and excessive decarburisation causes soft surface
hardness. Warping due to over heating is caused by excessive hardening strains.
Indication that hardening temperature may have been used are thumbnail cracks,
which are typical of tools with tapered sections, and deep cracks that indications of
too high a hardening temperature.
One indication that too low a hardening temperature has been used is that only the
outer portions of the tool have been heated and therefore only these portions have
gone through the transformation range. This nonuniformity can produce cracks during
quenching. Such cracking is characterised by “shelling off” of corners and edges (see
Figs. 3.3 and 3.4 for examples of improper hardening temperatures).

Fig. 3.3 Tool failures as the result of excessive hardening temperatures


(Bethlehem Steel Corporation)

1 2 3 4 5 6 7 8 9

Fig. 3.4 Tool failure as the result of too low a hardening temperature
(Bethlehem Steel Corporation)

Heat-treating instructions usually give a range of hardening temperatures and not just
one temperature. An experienced tool designer may call for the high side of the tem-
perature range when greater hardness penetration is required or the low side when
166 Tool Design

less hardness penetration or resistance to cracking is desired. The low side may also
be used when hardening small, long, thin parts that have a tendency to sag and flow
plastically at the hardening temperature. (This is not to say that they may not have to
have other support.)
Another factor concerning temperature is that the temperature controls heat-treating
equipment should be checked periodically. Faulty controls and indicating equipment
may cause too high or low a temperature without the heat-treater being aware of it.

Time Time in heat treating involves the rate at which the steel is heated to its hard-
ening temperature, the degree of soaking at this temperature, the length of soaking at
the tempering temperature, etc. Time is affected by various factors that will be pointed
out in the following discussion.
There has been considerable discussion about time required to heat tool steel to the
required hardening temperature. In the past, it was considered quite important to heat
the furnace and tool together in order to obtain a slow and uniform part, the reasoning
being that the surface of the steel was believed to reach the hardening temperature
before the inside. This has been proved untrue. According to literature published by
the Carpenter Steel Company, when the entire outside surface of a tool is up to fur-
nace heat, the centre is also up to heat. This is true regardless of mass. Knowledge
of this fact enables the centre as the steel will absorb the heat. It should be noted that
this does not apply to steels which have high hardening temperatures and require
preheating.
Thus, there are two methods used by hardeners whereby tools are heated for harden-
ing. One consists of heating the furnace and tool together, while the other consists
of placing the tool in a hot furnace that has already reached hardening temperature.
It is obvious that heating the tool and a furnace together would be time consuming,
especially in production treating, and therefore placing the tool in a hot furnace has a
definite advantage. The high quality of modern steel will allow this method of heating,
and the hardener need not be hesitant to use it. It may be necessary to heat the tools
and furnace together if the tool has considerable variation in section, as non-uniform
heating in this case could cause cracking.
The question now arises of how long the tool should be soaked after the tool has
reached the hardening temperature. It is desirable to use as short a soaking time as
possible in order to reduce decarburisation and grain growth.
An old rule of thumb has been to soak 1 hr per 25 mm of thickness of tool. Why so
long? If the inside of the tool comes to the same temperature as the outside almost
simultaneosly, certainly the transformation period will take place at about the same
speed. Soaking time need only be long enough to ensure that the entire tool has gone
through the transformation period. The real question is: When does the tool reach
hardening temperature?
Experienced hardeners are able to tell when the tool is up to heat by comparing its
colour with that of the thermocouple in the furnace. Some hardeners claim they can
tell by the “blush” on the tool. Whatever the method, it is recommended the hard-
ener use a soaking time of 5 min per 25 mm of thickness after the tool has reached
hardening temperature. Higher alloyed tool steels containing appreciable amounts of
chromium require longer soaking periods in order to obtain the proper carbide solu-
tion. The rule in this case is to soak at the hardening temperature 20 min plus 5 min
per 25 mm of thickness.
Tooling Materials and Heat Treatment 167

Time is as important in tempering as it is in hardening. When the tool is placed in the


tempering furnace, a certain period is required to reach the tempering temperature
and another period is necessary to provide soaking at the tempering temperature,
as was the case in hardening. Determining when the tool has reached the tempering
temperature is a problem, as there is no easy and practical way. Blacksmiths used
to determine tempering temperatures by oxide colours, but for the most part, this
method has been discontinued in modern toolmaking departments.
The type of tempering equipment will have an effect on the length of time required to
reach the tempering temperature. For example, a tool may reach tempering tempera-
ture in a bath or circulating air furnace in one-half the time it takes the tool to reach
temperature in a quiet hot-air oven.
Manufacturer’s instructions provide the hardener with tempering temperatures to
obtain the desired characteristics of a particular tool. They generally specify the
length of soaking time at the tempering temperature, meaning the tool should be
soaked this length of time after it reaches the tempering temperature. The total time
required to temper the tool is the given soaking time plus the time required to get up
to heat. Table 3.6 and 3.7 will give a general idea of the time needed to reach the
tempering temperature.

Table 3.6 Approximate time to reach drawing temperatures in a hot-air oven without
forced circulation
Drawing temperature °C Time, min per mm of thickness

Cubes or spheres* Long squares Average flat tools*


or cylinders*

149 1.18 2 3.5


204 1 1.8 2.6
260 1 1.6 2.4
371 0.8 1.4 2
482 0.8 1.2 1.6
*A form tool 76 mm diam by about 76 mm long would be “like a sphere.” Since it measures
76 mm thick, it would require about 75 min. to reach 204°C in a hot-air drawing oven without forced circulation.
A reamer would be “like a long cylinder”; and a blanking die measuring 102 × 32 × 204 mm would be an “aver-
age flat.”
SOURCE: Carpenter Steel Company, “Carpenter Matched Tool and Die Steel manual, “permission.

Vent
Atmosphere Before entering into discussion of
furnace atmosphere, it would be well to discuss the
various types of furnaces commonly used for heat
treating tools. Most furnaces for this type of heat
treating are muffle type, semimuffle type, liquid-bath Muffle
furnaces. A muffle furnace basically consists of a
closed refractory retort mounted in the approximate Burners
centre of the outer shell of the furnace. In gas-fired
furnace, combustion gases do not enter the retort
(muffle) but encircle it. The furnace atmosphere is Combustion area
ordinary air unless some means is provided to mod- Fig. 3.5 Cross section of gas-
ify it. Figure 3.5 shows a gas-fired muffle furnace. fired muffle furnace
168 Tool Design

Table 3.7 Approximate times to reach drawing temperatures in a circulating-air oven


or an oil bath
Drawing Times, min per mm of thickness

temperature, °C Cubes or Long squares or Average flat tools*


spheres* cylinders*

149 0.6 0.8 1.18


204 0.6 0.8 1.18
260+ 0.6 0.8 1.18
371 0.6 0.8 1.18
482 0.6 0.8 1.18
*A form tool 76 mm diam by about 76 mm long would be “like a sphere.” Since it measures
76 mm thick, it would require about 204°C, 75 min. to reach in a hot-air drawing oven without forced circulation.
A reamer would be “like a long cylinder”; and a blanking die measuring 102 × 32 × 204 mm would be an “aver-
age flat.”
†Oil baths should not be used above about 260°C.
SOURCE: Carpenter Steel Company, “Carpenter Matched Tool and Die Steel Manual,“ by permission.

Electric furnaces have the characteristics of


a muffle furnace in that heating elements are
placed around this muffle.
The semimuffle furnace consists of a refractory Refractory
hearth supported above the bottom of the outer
shell of the furnace. Gas burners direct the flame Burners
under the hearth, and the combustion products
rise through the space between the plate and
the sidewalls of the furnace and finally through Combustion area
vent holes in the furnace roof. The furnace atmo-
Fig. 3.6 Cross section of gas-fired
sphere consists of the combustion products, but
semimuffle furnace
the tools being heated do not lie in the direct
path of the flame. Figure 3.6 shows a semimuffle furnace.
Liquid-bath, or pot, furnaces consist of a pot containing a molten salt or metal sus-
pended into a refractory-lined furnace. It is usually fired by gas or oil or heated elec-
trically. There is no contamination from the outside atmosphere. Salts may provide
either a neutral or active bath. A neutral salt will heat the tool without any chemical
effects, and an active bath will impregnate the surface of the tools with carbon, nitro-
gen, etc., to give a hard surface.
Generally speaking, there are three types of atmosphere found in heat-treating fur-
naces: (i) neutral, (ii) oxidising, and (iii) reducing. A truly neutral furnace atmosphere
does not affect the composition of the tool being heated and is a term used rather
loosely among hardeners. A truly “neutral” atmosphere is hard to obtain. An oxidising
atmosphere contains an excess of oxygen caused by more air than necessary being
admitted to the fuel mixture (In gas-fired furnaces). A reducing atmosphere is just the
opposite. Not enough air is admitted to the fuel mixture to burn the fuel completely,
and as a result an excess of combustible gases is left in the furnace. Electric furnaces
are naturally oxidising because room air may enter the furnaces, unless some means
is provided to prevent air entry.
Tooling Materials and Heat Treatment 169

It has already been noted that the atmosphere surrounding the tool in a furnace has
a great deal to do with surface condition, along with the duration of heating and the
temperature level maintained. In order to control the atmosphere, many modern heat-
treating furnaces provide an artificial atmosphere whose characteristics produce the
desired results on the heat-treated part. These are known as controlled-atmosphere
furnaces. Some furnace atmospheres are produced which will protect the tool or heat-
treated part from oxidation or scaling, while others are formulated to prevent the addi-
tion of carbon in the surface of the part.
The principle of controlled-atmosphere furnaces is that gases are generated or pro-
duced outside the furnace and then piped into the heating chamber. Features are
incorporated into the furnace design to prevent loss of gas and to minimise have a
means of controlling the moisture content of the atmosphere. Figures 3.7 and 3.8
show one type of controlled-atmosphere furnace.

Fig. 3.7 Controlled-atmosphere box furnaces. Note atmosphere generator partially


hidden by left-hand furnace (Lind berg Hevi-Duty)

Since many tool and die shops still use furnaces that are not equipped with external
atmosphere controls, it would be well to discuss control of scaling and decarburisation
with reference to this type of furnace.
It should be mentioned at this point that the composition of various tool steels may
have a great deal to do with surface condition after heat treatment. Some types tend
to scale or decarburise more than others.
Contrary to common belief, an oxidising atmosphere does not necessarily mean a
decarburised surface. It will scale the surface, but in the case of many tool steels it will
not decarburise the metal underneath. On the other hand, a reducing atmosphere will
170 Tool Design

Fig. 3.8 Air-cooled 500-cfh Hyen endothermic atmosphere generator for use with box
furnaces (Lind berg Hevi-Duty)

not scale the surface, but if a large amount of water vapour is present, decarburisation
will be the result. Water vapour cannot be controlled on a manual furnace; therefore,
if a small amount of scaling is not objectionable, a manually controlled gas furnace
should be adjusted to excess oxygen to do a good hardening job. However, steels with
high hardening temperatures will have to be pack-hardened, as objectionable scaling
will occur at these higher temperatures.
Another method of preventing scaling and decarburisation is a carbon muffle. A muffle
of pure carbon is placed in the furnace, and tools to be heat treated are placed in the
muffle. The muffle generates its own atmosphere from its partial oxidation. This type
serves very well for hardening temperature in the 926 to 1316°C range (see Figs 3.9
and 3.10 for examples of a carbon muffle).

Fig. 3.9 The Sentry Diamond Block carbon muffle (Sentry Company)
Tooling Materials and Heat Treatment 171

Fig. 3.10 The Sentry Diamond Block carbon muffle in use (Sentry Company)

There are other device available commercially to help prevent scaling and decar-
burisation. One such device is a metal-foil container in which the tool is wrapped; he
whole container is placed in the furnace. The metal-foil container provides a protective
sheath and automatically neutralises the entrapped atmosphere (see Fig. 3.11).

Fig. 3.11 Sen/pak metal-foil containers used to provide protective sheath and automatically
neutralise the entrapped atmosphere (Sentry Company)
172 Tool Design

Cooling rate For a steel to harden it must be cooled fast enough to produce a fully
martensitic structure. The minimum cooling rate that will avoid the formation of softer
constituents other than martensite is known as the critical cooling rate. The critical
cooling rate is primarily determined by the composition of the steel.
The cooling rate determined by the quenching process is known as the actual cool-
ing rate. If the actual cooling rate is faster than the critical cooling rate, martensite
will form and full hardening of the steel will be the result. If the actual cooling rate is
less than the critical cooling rate, full hardening will not result, as softer constituents
(pearlite) will form along with martensite. Thus, it is important that the recommended
quenching medium be used for specific tool steels.
When a tool at hardening temperature is quenched in a liquid medium, the high tem-
perature of the metal will cause the medium to vapourise at the surface of the metal. A
thin film of vapour, sometimes referred to as a vapour blanket, will surround the metal.
Cooling at this stage is relatively slow, as a vapour film is a poor heat conductor. As
the metal cools, the vapour blanket dissipates and allows the medium to come into
contact with the metal. At this point violent boiling will occur, and the cooling rate is
increased. Further cooling reduces the temperature of the metal to the boiling point of
the medium, and thereafter cooling is by conduction through the liquid.
The most common quenching media are brine, water, oil, and air. Other media some-
times used are aqueous solutions of salts, acids, and alkalies. Brine has the fastest
cooling rate, with water, oil, and air following in descending order.
Fresh water is the most convenient of all the media, but it has the disadvantage of
forming a considerable vapour blanket on the surface of the work and vapour pock-
ets in holes and recesses. The result is uneven cooling, soft spots, and high cooling
stresses and may result in cracking of the work. In a still-bath-type quench, the work
should be moved through the medium with a slow up-and –down motion in order to
wipe off as much of the vapour blanket as possible. An even better method is to use
a flush-type quench, where the water is constantly circulated during the quenching
operation.
Brine is generally made by dissolving common rock salt in fresh water. A 5 to 10%
solution is often recommended. The addition of salt inhibits the formation of the vapour
blanket at the metal surface and allows direct contact between the bath and hot metal.
It is for this reason that the quenching process is completed faster in brine than in
fresh water and not because brine dissipates heat faster.
It should be noted that water is a satisfactory quenching medium for most purposes if
it is used as a circulated or flush quench. It is not as satisfactory as brine when used
as a still-bath quench.
The quenching rate of water drops as its temperature rises; therefore, sufficient vol-
ume should be provided to prevent too high a temperature rise. Cooling coils may be
inserted in the quenching tank to keep the medium at a constant temperature. Too low
a temperature may also be detrimental, since some steel are prone to crack in a cold
quench. Generally speaking, the temperature of water for quenching purposes should
be held between 15 and 32°C.
Oil quenching baths give a slower quenching speed than water and produce less
distortion and residual stress. They are well suited for complicated shapes and tool
Tooling Materials and Heat Treatment 173

steels that are difficult to harden without cracking. Mineral oil are generally used for
quenching because of their availability and lower cost. Vegetable and animal oils have
been used, but are not desirable as they become gummy from the heat of the metal
and produce offensive odours.
A good quenching oil should have a high flash point and a low viscosity. It should
be stable and chemically inactive with hot metal. It should not be used cold, as an
increase in bath temperature decreases the viscosity of the oil and causes a slightly
faster cooling rate. The temperature range recommended for an oil bath is from 38 to
54oC.
When air is used as a quenching medium, cooling depends upon radiation of heat
from the surface. The tool should be supported by means of a wire basket or screen in
order to permit free circulation of air on all sides. The tool should not be placed in cold
drafts. Intricate shapes may require a fan to circulate the air gently on all sides of the
tool, but a direct blast on one side of the tool should be avoided.
The tool being quenched in an oil or aqueous bath should not be dropped into the
bath because if the tool were allowed to rest on the bottom of the tank, one side would
cool faster than the other and could cause warping. Care should be taken not to strike
the tool on the side or bottom of the quench tank. Tools with thin or protruding sections
should be immersed in such a way that the bulkiest parts enter the bath first. Long
slender parts up-and –down agitation of the tool during the quench is preferred to a
swishing around movement, since the vertical movement permits more even cooling.
A heavy oxide scale may act as an insulator and retard the flow of heat from the steel
to the quenching medium. When a heavy scale has been allowed to build up and
minimum by proper hardening practise.
The tongs used to hold the parts for quenching should be designed to hold the part
securely and yet cover as small a portion of the tool as possible. Tongs that cover
large areas of the tool tend to cause uneven cooling. Some hardeners preheat tongs
to a dull red when hardening small part to ensure against locally chilling the part. In
regard to tongs, it is probably more important that they be designed correctly, for it is
quite easy to drop and ruin a part between the furnace and quench tank.
Tools should be tempered immediately after quenching, as residual stresses set-up
during quenching continue to build up and may cause the tool to crack if tempering
is delayed for a considerable length of time. If it is impossible to temper tools imme-
diately after quenching, tools may be held in boiling water, a steam bath, or a warm
bath of quenching oil.
A word of caution at this point. Tools may be tempered too soon. For example, if a
tool is pulled from the quenching bath before the quenching cycle is complete, full
hardening will not result because the actual hardening takes place at the end of the
quenching cycle, as explained earlier. A rule of thumb to follow is to quench the tool
down to a point where it can be held with the bare hands before beginning the temper-
ing operation.

3.11 HEAT TREATMENT AND TOOL DESIGN


Although the tool designer may not be required to perform actual heat-treating opera-
tions, he will design tools that require heat treatment. This section will show that faulty
174 Tool Design

design may cause cracking and distortion during heat treatment and discusses what
can be done in the way of design to prevent these tool failures.
Tools crack during heat treatment because of internal strains which exceed the
strength of the metal. The most serious internal strains develop during the quenching
cycle of hardening treatments and are usually the result of uneven cooling. The shape
of the tool determines the evenness of cooling. Heat storage may be greater in a
heavy section of the tool, and more sides may be exposed to the quenching medium
in other areas, as in the case of outside corners. Therefore, the tool should be shaped
to allow the heating and cooling rates to be as uniform as possible. This may be par-
tially accomplished by avoiding heavy section and sharp inside corners. Light sections
will cool rapidly and harden before adjacent heavy sections, and the result is exces-
sive quenching stresses. This type of trouble can be avoided by designing the tool as
a two-piece assembly or by using an air-hardening steel if a one-piece construction is
necessary. Figure 3.12 shows an example of this type of construction.

6 7 8 9 10

Fig. 3.12 Tool failure during heat treatment as the result of a light section adjacent to a
heavy section (Bethlehem Steel Corporation)

Sharp corners should be avoided in all tool designs, as they greatly concentrate
stresses in one small area. This is true not only in heat treatment but in service as
well. A good rule is to replace sharp corners with the largest filled possible. Many a
times trouble can be avoided by simply rounding inside corners. Keyways are a typical
example, as shown in Fig. 3.13.
Tooling Materials and Heat Treatment 175

Fig. 3.13 Tool failure through the corner of a keyway (Bethlehem Steel Corporation)

If sharp corners are necessary, an air-hardening steel should be used, since oil-and-
water-hardening steels are susceptible to cracking through sharp corners. Another
alternative is to fabricate the tool with a large fillet and then grind out the fillet after
hardening and tempering.
Holes are another source of trouble, as holes involve section changes which result in
the development of high stresses during heat treatment. This is particularly true when
hardness of the internal surface should be uniformly hard or uniformly soft. Flush
quenching may be required to harden small holes effectively.
Uneven quenching can be avoided in holes by packing them with a suitable material
before hardening, for example, fireclay, steel wool, or asbestos. Steel wool is one of
the best materials for this purpose, as it is easily removed after heat treating.
Blind holes should be avoided whenever possible. They are especially troublesome
if they are required to be uniformly hard. In this case, an air-hardening steel should
be used. Blind holes may be packed with steel wool if they are not required to be
hardened internally.
Threaded holes in a heat-treated tool are also objectionable. The treads act as stress
risers and frequently cause cracking during a liquid quench. Threads should be placed
in the softer components of the tool whenever possible, with a thorough counterbored
hole in the hardened component to accept the cap screws (see Fig. 3.14).

Hardened tool
component
Dowel pins Soft tool
Socket-head component
cap screw

Fig. 3.14 Toll components showing threads in soft component


176 Tool Design

The location of holes and hole spacing in heat-treated tools may have an effect on
stresses during heat treatment. Holes near the edge of the tool or close together
have the same effect as thick sections adjacent to thin sections. A general rule for
hole location is to allow at least one hole diameter between the edge of the tool and
the hole. This distance should be 1.5 times the hole diameter for holes located in
the corners of the tool. Adjacent holes should have at least the thickness of the tool
between them. If the above rules cannot be adhered to, the tool should be made from
an air-hardening steel (see Fig. 3.15).

1 2 3 4 5 6 7

Fig. 3.15 Tool failure during heat treatment because of holes too close to the edge of the
tool (Bethlehem Steel Corporation)

The tool designer should also be aware that fabrication flaws may also result in tool
failure during heat treatment. Deep scratches
and toolmarks may serve as stress concentra-
tion points. Rough holes, drill marks, countersink
marks, and deep centre holes may be sources
of trouble.
Improper grinding of hardened and tempered
tools may result in tool failure. Grinding involves
intense localised heating at the point of abra-
sion, which may set up high surface stresses in
an already stressed tool. The result may be sur-
face cracks, which appear in a network pattern
characteristic of grinding cracks. Careful grind-
ing techniques, such as frequent wheel dress-
ing, effective use of coolant, softer wheel grades,
lighter feeds and depths of cut, coarser grinding
wheels, etc., will prevent grinding cracks (see 1
2 "
Fig. 3.16). 8

The use of steel letters to stamp tooling identifica- Fig. 3.16 Network pattern of
tion on tools which are to be heat-treated should grinding cracks on heat-
be avoided. The stamp marks provide stress ris- treated tool (Bethlehem
ers which cause concentration of stresses in one Steel Corporation)
Tooling Materials and Heat Treatment 177

area. If identification must be stamped, avoid deep stamp marks. Other methods of
marking should be used when possible (see Fig. 3.17).

Fig. 3.17 Tool failure during heat treatment because of deep stamp or tool-identification
marks (Bethlehem Steel Corporation)

The various types of tool failures which have been discussed take place during heat-
treating operations. If the heat-treating operation could be eliminated, or if heat treat-
ment could be done to a solid block before machining intricate tool shapes, these
failures could be eliminated. This is possible in some cases if machining processes
such as electrodischarge machining or electrochemical machining are available. Tool-
steel companies offer prehardened tool steel in various hardened conditions for this
purpose.

Summary

Tool must be hard, wear resistant, and have very low surface roughness. Conventional
materials do not cater to these needs; as such tooling requires special materials. This
chapter aimed to acquaint the readers with the various available tool materials.
• Tool steels can be used for tooling applications requiring high strength, hardness, toughness,
and wear resistance.
• Water hardening tool steels are recommended for light or medium cold impact operations such
as coining, cold heading, punching, knurling, embossing, etc. The water hardening tool steels
should not be used when tool has drastic dimensional changes.
• Shock resisting tool steels are recommended for pneumatic chisel, heavy duty shear blades,
punches, rivet busters, and similar tools.
• Oil hardening tool steels should be applied for applications such as blanking, bending, trimming,
coining, shearing and shaping dies.
• Air hardening tool steels are applicable for intricate tool shapes such as thread rolling dies, long
slender broaches etc.
• High carbon high chromium cold work tool steels are suitable for wire drawing dies, master
gauges, blanking and piercing dies, etc.
• Hot work tool steels are used for applications such as hot forging dies, hot extrusion dies, hot
shears, die casting dies, and plastic moulding dies.
• High speed tool steel are mainly used for metal cutting applications, although they can also be
used as extrusion dies and blanking punches and dies.
178 Tool Design

• Cast iron can be used as large forming dies and drawing dies and as material for die set
shoes.
• Mild or low carbon steel are also used extensively in tool making. One of the principle applica-
tions of this material is for jigs and fixture.
• Nonmetalic tooling materials include rubber, masonite, densified wood, plastics, oxide cutting
tool materials, diamond, etc.
• Nonferrous tooling materials include sintered carbide, cast nonferrous alloys, bismuth alloys,
magnesium, etc.
• Heat treatment is a process whereby the physical properties of the material can be change
suitably by subjecting it to a combination of heating and cooling. Heat treatment processes are
widely used techniques to achieve the required properties of the tooling material. Various heat
treatment processes includes, normalising, annealing, spheroidising, stress relieving, stabilis-
ing, hardening, pack hardening, quenching, tempering, decarburising, etc.

Questions
1. What type of load is imposed upon the work material during a punching operation?
2. What is a ductile material?
3. Why are the properties of elasticity and stiffness important to the designer of cutting tools?
4. Give an example of a tool designer’s product that must be made of material exhibiting the prop-
erty of toughness.
5. What is meant by red hardness?
6. What is the Rockwell hardness of a metal-cutting file?
7. What tool steels are harder than a metal-cutting file in a heat-treated condition?
8. The machinability rating of B1112 screw stock is 100. The machinability rating of group W tool
steels is also 100. Does this mean the adaptability to cutting is the same for both materials?
Why?
9. What is meant by the term tool steel?
10. What are the major elements in group W tool steels? What effect does each element have on
the group W steels?
11. Are group W steels suitable for hot-work applications? Why?
12. Why do group W steels machined to complicated shapes have a tendency to crack or warp dur-
ing quenching operations?
13. What are the chief alloying elements of group S tool steels?
14. What are the major advantages of group O steels compared to group W steels?
15. What is the major advantage of group A tool steels?
16. What are the characteristic of group D steels?
17. What is the major characteristic of group H tool steels?
18. What is the primary use of high-speed tool steels?
19. What are the two groups of high-speed steels, and how are they designated?
20. What effect does the addition of cobalt have on high-speed steels?
21. What is the advantage of using gray cast iron as the material for the main body of a jig or
fixture?
22. Why is hot-rolled carbon steel suited for tools that are to be fabricated by welding?
Tooling Materials and Heat Treatment 179

23. What determines the choice between plastics and conventional tooling material for tooling
applications?
24. What is the basic component of oxide cutting tools?
25. What is the chief limitation of oxide cutting tools?
26. What are the characteristics of machine tools utilising oxides as cutting tools?
27. What are the advantages of diamonds as a production turning tool?
28. What are the two basic groups of tungsten carbide cutting tools?
29. What are the major elements in cast nonferrous cutting tools?
30. Compared to carbide and high-speed-steel cutting tools, what are the major advantages of cast-
alloy cutting tools?
31. What type of grinding equipment is necessary for sharpening cast-alloy cutting tools?
32. Why are high-bismuth alloys excellent materials for duplicating mould detail?
33. Why is magnesium often used as structural material for large assembly fixtures in the aircraft
industry?
34. What is meant by the general term heat treating?
35. How is normalising accomplished?
36. Generally speaking, what is the purpose of annealing?
37. How does spheroidising increase the machinability of a tool steel?
38. Why is the critical temperature important to the heat treater?
39. What is the major difference between stress relieving and stabilising?
40. What is pack hardening?
41. What is the difference between tempering and drawing?
42. What is the purpose of tempering?
43. Why is it necessary to bark the surface of tool steels before they are heat treated?
44. Why is it necessary to preheat some of the highly alloyed tool steels?
45. In what from does carbon occur in steel?
46. What is the metallurgical term for iron carbide?
47. What is pearlite?
48. When steel is heated, what physical changes occur as the temperature passes 720°C?
49. What is the critical range, or transformation range, of a given steel?
50. What effect does carbon have on the transformation range?
51. What changes take place in the steel upon further heating above the transformation range?
52. What happens when steel is allowed to cool slowly through the transformation range?
53. What determines the final effective grain size after heat treatment?
54. Why is knowledge of the change of volume that occurs during transformation important to the
heat treater?
55. What happens when tool steel is suddenly cooled through the transformation range?
56. Why is it necessary to temper tool steel in the as-quenched martensitic condition?
57. What are the indications that a tool has been subjected to too high a hardening temperature?
58. What are the indications that the tool has not reached the proper hardening temperature?
59. How long a tool should be soaked after it has reached the proper hardening temperature?
60. What are the major advantages of liquid-bath or pot-type heat-treating furnaces?
180 Tool Design

61. What are three types of atmosphere found in heat-treating furnaces? What are their
characteristics?
62. Name the various methods of controlling scaling and decarbonisation of workpiece during heat
treatment?
63. What is the critical cooling rate of tool steel?
64. What are the most common quenching media in use today in decreasing order of quenching
severity?
65. How is brine made?
66. What is the best temperature of water used for quenching purposes? Why?
67. What are the characteristics of good quenching oil?
68. Why should be tools tempered immediately after quenching?
69. Why is a tool component that is constructed with a heavy section adjacent to a thin section likely
to crack during heat treatment?
70. What should be done when it is necessary to make a tool component containing sharp corners,
holes near the edge, blind holes, or thick sections adjacent to this sections?

Problems
1. A die block with the dimensions of 25 × 100 × 100 mm is to be made from O2 tool steel. Can the
toolmaker successfully construct the die block from a 25 × 100 mm of O2 tool steel?
2. Six punches whose dimensions are 20 mm diameter by 75 mm long have been finish-machined
and are ready for heat treatment. The punches are made from O2 tool steel. (a) How long they
should be soaked in the hardening furnace before quenching? (b) What is the approximate time
required to reach a drawing temperature of 200°C?
3. Study the O2 die block shown in Fig. 3.18. Give five reasons why it might fail by cracking during
heat treatment.
Identification
number stamped
into die block

2076

Die Die
opening 1 opening 2

Dowel-pin C-bore holes


holes (2 each) (4 each)

Fig. 3.18 O2 die block: will it fail during heat treatment?

4. Select the proper tool steel for the following applications and list the properties that make it
applicable: (a) shear blade for cold work (12 mm plate), (b) hand reamer (solid), (c) ring gauge,
(d) end mill, (e) punch pilot, (f) wire-drawing dies (cold).
Tooling Materials and Heat Treatment 181

References
Books
• ASM: “Metals Handbook,” 8th ed., Metals Park, Ohio; “Properties and Selection of Metals,”
1964; “Heat Treating, Cleaning, and Finishing,” 1964; “Machining,” 1967.
• ASTME: “Tool Engineers Handbook,” 2d., F. W. Wilson (ed.), McGraw-Hill, New York, 1959.
• Avner, S H: “An Introduction to Physical Metallurgy,” McGraw-Hill, New York, 1964.
• Bethlehem Steel Company: “The Tool Steel Trouble-shooter,” Bethlehem, Pa., 1952.
• Carpenter Steel Company: “Carpenter Matched Tool and Die Steel Manual,” Reading, Pa.
• Palmer, F R, and G V Luerssen: “Tool Steel Simplified,” 3d ed., The Carpenter Steel Company,
Reading, Pa., 1964.
• Roberts, G A, J C Hamaker, and A R Johnson: “Tool Steels,” 3d ed., American Society for
Metals, Metals Park, Ohio, 1962.
Catalogues
• Allegheny Ludlum Steel Corp., Pittsburgh, Pa.
• Bethlehem Steel Company, Bethlehem, Pa.
• Carpenter Steel Company, Reading, Pa.
• Crucible Steel Co. of America, Pittsburgh, Pa.
• Peninsular Steel Co., Detroit.
• Timker Roller Bearing Company, Canton, Ohio.
4 DESIGN OF
CUTTING TOOLS

4.1 INTRODUCTION
The most common method of metal removal is to use an edged cutting tool. This pro-
cess is based upon the separation of the base metal by pressure applied with a cut-
ting tool made from a harder material. It has developed through the ages from a crude
chisel-like object powered and guided by the human hand to the present state of the
art, where a powerful machine tool forces and automatically guides a multiedged cut-
ting tool through a large block of metal.
The basic elements of this modern metal-removal process consist of a machine tool,
a control system, and the cutting tool. Each element may be compared to the leg of
a tripod. If one leg is missing, the other two will fall. Thus, a machine tool and control
system are useless without cutting tools and vice versa. This chapter will be con-
cerned with the cutting tool “tripod leg” because one of the primary duties of the tool
designer is to select or design tools for a metal-cutting operation.
Fortunately, the tool designer’s job has been simplified in recent years because cut-
ting-tool manufactures have made standard tools available in a wide variety of shapes
and sizes. Most machining operations can be performed by the application of standard
cutters. In this case, the major problem of the tool designer is to select the standard
cutting tool which will do the job in the most efficient and economical way. To select the
proper tool effectively, he must have a good knowledge of the metal-cutting process,
be familiar with tool geometry, and be aware of the types of standard tools available.
It should be emphasised that standard cutting tools should be used whenever pos-
sible for reasons of economy. The design and manufacture of special cutters may be
necessary in rare instances, but this practise should be avoided whenever possible.
Cutting-tool manufactures mass produces their products and are able to keep the
cost to a minimum. One special tool manufactured in the designer’s own plant may
cost, many-a-time, the amount of a standard tool. An added advantage in using stan-
dard tools is that they can readily be replaced.

4.2 A BRIEF HISTORY OF METAL CUTTING


The first metal cutting probably occurred during the Bronze Age; however, the first
real evidence of progress in metal cutting was found after the discovery of iron. It is
believed that man discovered iron nearly 6000 years ago. The first iron was prob-
ably extracted from meteorites, which ancient people often referred to as “metal from
heaven.” Iron was considered a rare and precious article, as evidenced by the iron
relics and jewellery found in Egyptian and Roman cemeteries and ruins.
The earliest written reference to iron manufacture is in Genesis, which records that
Tubol-cain, seven generations from Adam, was a master of coppersmiths and black-
smiths. Iron articles then were probably forged, chiselled, or filed. There is also evi-
dence that a cutting tool in the form of a chisel-point drill was used for the drilling holes
Design of Cutting Tools 183

in brass and iron in ancient Egypt. The drill, therefore, is one of the oldest cutting tools
known to man.
The first real step in metal cutting was taken in the latter part of the eighteenth cen-
tury. In 1775, Wilkinson invented his boring machine which enabled Watt to carry on
the development and construction of the steam engine. Following this achievement,
Mardslay developed his practical screw-cutting lathe in 1880. The first drill with auto-
matic power feed was developed in 1840 by Namsmyth. Richard Roberts developed
the planer is 1817. Eli Whitney developed the first milling machine in 1818, using an
improved version of the rotary-file principle previously used by Swiss watchmakers.
This was the great era for the development of the basic metal-cutting machine tools
as we know them today.
At the turn of the nineteenth century, metal-cutting tools were made from high-carbon
steel. This material took a very keen cutting edge and is still used in the manufacture
of files, hand reamers, wood-cutting tools, and, in some instances, tools for machin-
ing aluminium and magnesium. However, the temperature of the cutting edge must
be held below 200°C if the tool is to hold its cutting edge. In the early 1900s, Frederic
Taylor and associated metallurgists developed high-speed steel and very nearly revo-
lutionised the whole metal cutting industry. Cutting tools made of this material retain
their cutting edge at temperatures of over 550°C. Speeds at which metal could be
machined were more than doubled as a result.
The next significant development in metal cutting came in the twenties, with the intro-
duction of sintered carbides as a cutting tool. Again cutting speeds were doubled. The
latest significant development has been the use of oxide (ceramic) cutting tools, and
carbide cutting speeds were doubled. These advances in cutting-tool materials have
necessitated the redesign and development of heavier and more productive machine
tools.
From this discussion, it can be seen that metal cutting has developed over a period
of centuries. The development was slow in the beginning, and the greatest advances
have been in the past 150 years. What will metal-cutting concepts be like if they are
carried on into the future with more accelerated progress than in the past? Whatever
the outcome, it should be noted that the present objectives of tool-material develop-
ments are effective hardness at higher temperatures, improved impact and shock
resistance, and improved wear characteristics. Each new development in cutting tools
advances this goal still further, although in some cases one objective may be sacri-
ficed to another advances. For this reason, each tool material has tended to supple-
ment the older materials. There have been very few cases where a new cutting-tool
material has completely replaced an older material.

4.3 METAL-CUTTING PROCESS


4.3.1 Basic Requirements of a Cutting Tool
The first requirement of all metal-cutting tools is that the tool must be harder than the
material being cut. The second is that the tool and material being cut must be held
rigidly in a manner that will cause the tool to penetrate the workpiece when forces
are applied. When these conditions are satisfied, the metal will be displaced, provid-
ing there is sufficient force and power to overcome the resistance of the workpiece
material.
184 Tool Design

Figure 4.1 shows the progressive formation of a chip using a single-point tool. At
a, the tool contacts the workpiece material. Compression in the workpiece material
takes place at the point of contact at b. At c, the cutting force overcomes the resis-
tance to penetration of the tool and begins to deform by plastic flow. As the cutting
force increases, either a rupture or a plastic flow in a direction generally perpendicular
to the face of the tool occurs, and the chip is formed as shown at d. The formation of a
new chip segment begins, and the sequence of compression, plastic flow, and rupture
is repeated. The cycle is repeated indefinitely. This process of chip formation is basic
to edged metal-cutting tools, and will be discussed in more detail in Section. 4.4.

Cutting Cutting
Tool force Tool
force

Workpiece
material

(a) (b)

Cutting Cutting
force Tool force Tool

(c) (d)

Fig. 4.1 Progressive formation of a metal chip


Rake angle
Consider for a while the shape of the cutting
tool. Figure 4.2 shows a tool similar to the one in
Fig. 4.1, except that the general shape is defined
in two dimensions. The terms commonly used to
define the shape of the tool are rake angles and
relief angles. Cutting
tool
Rake angles vary from positive to negative (see
Fig. 4.3). Generally speaking, positive rake angles
have greater cutting efficiency from the stand-
point of power requirements and cutting forces. Relief angle
The smaller point angle resulting from positive
rake angles penetrates the metal more readily
and reduces cutting pressures. It should be noted,
however, that high positive rake angles result in a
Fig. 4.2 Rake and relief angles of
fragile cutting edge. They are limited to machining
a cutting tool
Design of Cutting Tools 185

Fig. 4.3 Variation of rake angles from positive to negative: (a) positive rake (b) zero
rake (c) negative rake

softer materials and are well suited to this type of application because of the keenness
of the cutting edge. On the other hand, negative rake angles provide a stronger cutting
edge and are suitable for cutting high-strength alloys. Relief angles, sometimes incor-
rectly referred to as clearance angles, are to allow the cutting edge to penetrate the
work piece (see Fig. 4.4). They prevent rubbing or physical interference between the
heel of the tool and the workpiece.
Relief angles are always positive. An edged cutting tool must always have positive
relief to cut successfully. Small relief angles give more support to the cutting edge and
are generally used when machining high-strength alloys. Larger relief angles provide
a keener cutting edge and allow the tool point to penetrate the work material more
easily. However, as in the case of high positive rake angles, a large relief angle results
in a fragile cutting edge that is suitable only for softer low-strength alloys.
Relief and rake angles are common to all edged cutting tools, although there may be
small differences in terminology when referring to different tools. For example, in the
common twist drill (Fig. 4.5) the helix angle is equivalent to the rake angle of a single-
point tool. The lip relief angle is equivalent to the relief angle of a single-point tool.
Another example is shown in Fig. 4.6. The relief shown on the mill cutter is referred
to as end relief, while the rake angle is referred to as axial rake. Table 4.1 shows the
approximate equivalents of relief and rake angles for common edged cutting tools.
Thus, the tool geometry of edged metal-cutting tools is basically the same. Generally
speaking, tool geometry is a matter of rake and relief angles, although different
186 Tool Design

Small relief angles


provide stronger
cutting edge

Workpiece

Zero or negative relief


allows heel of tool to drag
on workpiece material and
prevent tool from entering

Workpiece

Greater relief angles provide a keener


cutting edge that will penetrate the
work piece more readily but also
give a fragile cutting edge

Workpiece

Fig. 4.4 Effect of variation in relief angles

Helix angle

Axial rake

Lip relief
angle
End relief

Fig. 4.5 Relief- and rake-angle equiv- Fig. 4.6 Relief- and rake-angle equivalents
alents of a twist drill of a face-milling cutter

terminology is used on different tools. However, this is not to imply that tool geometry
is simple. A great deal of research on the subject is still being carried out. Different
cutting-tool materials require different tool geometry, and additional research will be
necessary as new cutting-tool materials are developed.
Design of Cutting Tools 187

Table 4.1 Approximate equivalents of tool geometry


Lathe and planer Milling cutter Drill Broach

Back rake Axial rake Helix angle Hook angle


Side rake Radial rake Radial rake Side rake
Side cutting-edge Corner angle Half-point angle
angle
End cutting-edge angle End cutting-edge angle
Side relief Peripheral relief Clearance
End relief End relief Lip relief Clearance

SOURCE: Metcut Research Associates, “Machining Data Handbook,” 1966, by permission.

4.4 MECHANICS AND GEOMETRY OF


CHIP FORMATION
Until recent years, developments in metal cutting evolved through trial-and-error meth-
ods. The use of metal-cutting tools was learned on an empirical basis without really
understanding the basic principles involved in the metal-cutting process. Progress
was made, but the rate of progress was slow. In order to meet the demands of modern
industrial technology, it became necessary to accelerate this progress.
In recent years, certain men concerned with machine tools have recognised this basic
fact and have endeavoured to obtain a clear understanding of what is taking place
at the cutting edge. They have attempted to deal with the quantitative as well as the
qualitative aspects, and have succeeded. They speak in terms of forces, stresses,
strains, coefficients of friction, velocities, required power, etc. Knowledge of this kind
makes greater metal-cutting efficiency possible, which in turn results in the greater
production output demanded by modern industrial technology.
The average tool designer usually does not carry on formal metal-cutting research
as part of his regular job. His main concern is to apply the results of formal research
to his particular situation. For example, he is more concerned with applying a forces
formula that has been simplified and generalised that with the derivation of that for-
mula. He wants to know the correct size of tool shank for his machining application
immediately and is not concerned with the mechanics involved in the development
of the formula. The results of research and development departments of the major
metal-cutting tool industries are usually compiled in the form of simplified formulas,
graphs, and nomographs designed to fit the production man’s needs. The average
tool designer’s responsibility lies in knowing where to obtain this information, as well
as how to apply it. The information given in this chapter deals with this aspect of
tool design. A rigorous mathematical treatment of metal-cutting physics is beyond
the scope of this text. The purpose of the information given in this chapter is to give
the prospective tool designer background information that will enable him to apply the
results of metal-cutting research.1

1
If he is so inclined, he should investigate the works of Boston, Merchant, Ernst, Kronenberg, Trigger, and
others who have contributed to the formal study of what really happens at the cutting edge.
188 Tool Design

Chip formation The basic formation of a metal chip was discussed earlier in the
chapter. Chip formation is shown photographically in Fig. 4.7. Generally speaking, it
may be said that this is the basic mechanism by which all types of chips are formed
in cutting metal. There is always a deformation of metal lying ahead of the cutting
edge by a process of shear. Figure 4.8 illustrates this mechanism in general terms.
Here the metal deforms by shear in a narrow zone extending from the cutting edge
to the work surface. This shear zone is treated as a single plane for the purposes of
mathematical analysis and is commonly referred to as the shear plane. The angle
which the shear plane makes with the direction of travel of the tool is known as the
shear angle.

Fig. 4.7 Formation of a chip by a metal-cutting operation


(National Twist Drill and Tool Company)

The process of plastic deformation occurring along the plane elongates the individual
crystals of the metal in the general direction indicated by the shear angle. This action
tends to produce a chip that is thicker than the layer of parent material from which it
came. Referring again to Fig. 4.8, note that the chip material moves up the tool face in
layers of distorted material. Each layer is pushed outward a fixed amount with respect
to an adjacent layer and retains this position as the whole chip slides up the tool face.
The distorted layers flow by means of a phenomenon referred to as slip and the layers
are called slip planes. The number of slip planes depends upon the lattice structure of
the parent workpiece material. The distortion of the layers tends to strengthen them
(work hardening), and therefore the hardness of the chip is usually much greater than
the hardness of the parent material.

Distorted metal
Plan
e of Tool
shea
r
Shear angle

Workpiece

Fig. 4.8 The basic mechanism by which all chips are formed
Design of Cutting Tools 189

Thus, in simple language, the mechanics of chip formation in any machining opera-
tion is a rapid series of plastic flow and slip movements ahead of the cutting edge.
This produces a vibration of high frequency at the tool tip, which must be absorbed
or dampened by the cutting edge and its support. Excessive overhang of a tool or a
poorly supported workpiece aggravates this vibration of the tool, and the result may
be chatter marks on the finished surface and reduced tool life.
The degree of plastic flow ahead of the cutting tool determines the type of chip that
will be produced. If the workpiece materials brittle and has little capacity for defor-
mation before fracture, the chip will separate along the shear plane to form what is
known as a discontinuous or segmental chip. Materials that are more ductile and
have the capacity for plastic flow will deform along the shear plane without rupture.
The planes tend to slip and weld to successive shear planes, and the result is a chip
that tends to flow in a continuous ribbon along the face of the tool. This is known as
a continuous chip and is usually much harder than the parent material because of its
strain-hardened condition.

Types of chips Generally, there are three basic types of chips. The discontinuous
and continuous have already been mentioned. The third is the continuous chip with a
built-up edge (BUE). The “Tool Engineers Handbook” lists a fourth type, the inhomo-
geneous chip. The study and investigation of chip types is helpful in understanding the
metal-cutting process. Therefore, each type will be given individual consideration.
1. Discontinuous chips This type is gener-
ally formed as the result of machining a brittle
material, such as grey cast iron and high-lead
bronze. A thick chip in some materials will peri-
odically rupture along the plane of shear. The
chip may come off in completely individual seg-
ments, or the segments may adhere to each
other to form a loose chip which fractures eas-
ily. The main advantage of the chip is that it Fig. 4.9 Formation of a
is easily disposed of during production runs. discontinuous chip
The resulting surface finish may be fair to good
when the chip is produced from the more brittle materials. Figure 4.9 shows the for-
mation of a discontinuous chips (see Fig. 4.11 for a photograph of a discontinuous
chip).
2. Continuous chips These are associated with relatively ductile materials where defi-
nite successive ruptures do not take place. The planes tend to slip and weld and pro-
duce a continuous plastic flow of the metal without fracture ahead of the tool, followed
by a smooth flow up the face of the tool. This type is usually associated with higher
cutting speeds, as low cutting speeds sometimes cause certain ductile materials to
pile up ahead of higher cutting speeds, as low cutting speeds sometimes cause cer-
tain ductile materials to pile up ahead of the tool and come off in small lumps similar
to discontinuous chips. It is often referred to as an “ideal” chip because it produces
excellent surface finish, the power consumption is low, friction between chip and tool
is low, and excessive heat is not generated. It does create a chip-disposal problem,
and this is usually taken care of by incorporating chip breakers behind the cutting-tool
edge. Figure 4.10 shows the formation of a continuous chip, and Fig. 4.11 shows a
photograph of one.
190 Tool Design

3. BUE chip Although continuous chips are Tool


sometimes considered ideal, they are often travel
difficult to obtain when machining the usual Tool
ductile materials. More often than not, ductile
materials appear to be “gummy” and tend to Smooth surface
adhere to the face and flank of the tool. This
is the result of high friction between the face
of the tool and the chip. The metal piles up
on the tool point until it appears to resemble Fig. 4.10 Formation of a continuous
a cutting edge and is sometimes incorrectly chip
referred to as a “false cutting edge”. More cor-
rectly, it is referred to as a built-up edge (BUE), and the result is a continuous chip
with a built-up edge.
BUE chip is common to most metal-cutting operations (Figs 4.11 and 4.12). It is asso-
ciated with ductile materials being machined at speeds in the high-speed-steel cutting
range (usually below the carbide cutting range). The built up edge is a glob of work
material that adheres to the face of the tool as the chip shears past it. It is the result
of high resistance to the sliding of the chip up the tool face. This high friction causes
some of the chip metal to shear away from the body of the chip and remain more
or less stagnant on the tool face. Once the BUE is established, metal is continually
sloughed off on the passing chip and work surface. New material is continually added
to the BUE but in turn continues to pass off, as described above. The result is poor
surface finish because fragments are continually shed on the finished surface. In the
case of carbide tools, small particles of the cutting edge may be removed as the
BUE sloughs off, resulting in a minute chip. Repeated chipping may cause premature
failure of the cutting edge. This condition is usually prevalent at extremely low cutting
speeds.

(a) (b) (c)

Fig. 4.11 Basic chip types: (a) Discontinuous (b) Continuous (c) Built-up edge
(Cincinnati Milacron, Inc.)

Generally speaking, any change in cutting conditions that will reduce or eliminate
the BUE is desirable. Since the major cause of the BUE is high friction between the
chip and tool face, any means of reducing this friction should also reduce the BUE.
The application of a lubricant and adhesion-preventing agent is often quite effective,
especially when it is necessary to operate at low-cutting speeds. A high polish on the
Design of Cutting Tools 191

tool face will also help to reduce friction. The use of a tool material having an inherent
low coefficient of friction combined with the ability to withstand high-cutting speeds
may be desirable for high production. The steel-cutting grades of carbide and oxide
cutting-tool materials are examples for this type of application.

Tool
travel Tool

BUE

Rough surface

Fig. 4.12 Formation of a continuous chip with a built-up edge

4. Inhomogeneous chip The chip1 is produced by a process similar to necking in


a tensile specimen. The shear deformation occurring in chip formation causes the
temperature of the shear plane to rise, which in turn may decrease the strength of
the material and cause further strain if the metal is a poor conductor. This process,
when repeated several times, results in a large strain at the point of initial strain. Then
a new shear plane will develop some distance from the first, and the deformation
shifts to this point. The resulting chip is banded with regions of large and small strain
as shown in Fig. 4.13. This is characteristic of metals suffering a marked decrease in
yield strength with temperature and poor thermal conductivity. Titanium-alloy chips
frequently are of this type.

Tool
travel
Tool

Fig. 4.13 Formation of an inhomogeneous chip

In summary and generally speaking, the more brittle materials tend to produce dis-
continuous chips, while the more ductile materials tend to produce continuous chips.
It is possible, however, to encounter more than one type of chip with one material.
Under different operating conditions of cutting speed, chip thickness, lubrication, tool
geometry, etc., it is possible, and many a times desirable, to change from one type of
chip to another, even though the work material remains the same.

Orthogonal and oblique cutting For experimental purposes, metal cutting has
been classed as either orthogonal or oblique. The simplest is the orthogonal cutting

1
Described in ASTME, “Tool Engineers Handbook,” 2d ed., F W Wilson (ed.), McGraw-Hill, New York, 1959.
192 Tool Design

process. Here the tool is set with its cutting edge perpendicular to the direction of tool
travel, and only a straight cutting edge is active. Experimentally, this type of cutting
is accomplished by planning the edge of a plate with a tool wider than the thickness
of the plate. Comparable metal-cutting operations in industry are lathe cut-off opera-
tions, grooving operations, certain types of broaching and plane-milling operations
utilising a straight tooth plain mill.1 The primary advantage of orthogonal cutting is
the two-dimensional nature of the patterns of deformations and stresses. The many
variables involved in metal cutting can thus be studied under the simplest geometrical
configuration. Its main use is to achieve the basic foundations and analytical tech-
niques for treating more complicated cases.
Unfortunately, the simple geometry of orthogonal cutting does not apply to the major
portion of metal cutting in industry. Oblique cutting is a better example of metal cut-
ting in industry, although the force geometry is more complicated. In oblique cutting,
as opposed to orthogonal cutting, the single straight cutting edge is inclined with
respect to the direction of tool travel. In other words, positive or negative side rake is
added to the tool. Thus, the oblique cutting process is three-dimensional. The majority
of industrial metal-cutting operations are oblique; examples are drilling, conventional
turning, face milling, and other operations where the tools utilise a positive or negative
side-rake angle (or equivalent).
Analysis of metal cutting based on oblique cutting is complex. Therefore, the pre-
ferred approach is orthogonal cutting for metal cutting analysis. Orthogonal cutting is
only an approximation of the actual scenario, and the assumptions used in orthogonal
metal cutting analysis are noted below:
1. No contact at the flank, i.e., the tool is perfectly sharp.
2. No side flow of chips, i.e., width of the chips remains constant.
3. The cutting velocity remains uniform.
4. A continuous chip is produced with no built-up edge.
5. The work material undergoes deformation across a thin shear plane.
6. There is uniform distribution of normal and shear on the shear plane.
7. The work material is rigid and perfectly plastic.
Figure 4.14 shows an example of orthogonal cutting. Here the importance of the
shear angle is illustrated. The shear angle is determined by the rake of the cutting
tool, type of material being cut, coefficient of friction of workpiece material and tool
material, type of lubricants being used, etc. In practise it is variable in size, and every
effort should be made to keep it as large as possible. The reason is that when the
angle of shear is small, the path of the shear will be longer and the resulting chip will
be thicker and will require greater cutting forces to separate the chip from the work-
piece material. The reverse is true if the shear angle is large.
In order to determine the shear angle mathematically, it is first necessary to determine
the cutting ratio or chip thickness ratio. The cutting ratio or chip thickness ratio, r, is
defined as the ratio of uncut chip thickness t0 and chip thickness tc (Refer to Fig. 4.14).
Thus, from geometry.

1
The student should note the distinction: plane milling refers to a flat surface with reference to which the
workpiece is located; plain milling refers to the simplest type of machine or accessory, such as a plain milling
cutter or plain indexing.
Design of Cutting Tools 193

Fig. 4.14 Orthogonal cutting showing the effect of large and small shear angles on chip
thickness and length of shear planes

t0 sin f
r = __ = _________ (4.1)
tc cos (f – a)
The chip thickness ratio is sometimes used as a measure of the quality of the sur-
face finish. When r is high, the cutting action and quality of the surface are generally
good.
Eq. 4.1 can be rearranged to have a relationship of shear angle with clip thickness
ratio and rake angle.

r cos a
tan f = _________ (4.2)
1 – r sin a

Velocity relationships If Vc is the cutting


velocity, Vf is the chip flow velocity and Vs
is the shear velocity, a vector triangle can
be formed in the cutting zone in order to
establish a relationship of Vr and Vs with
Vc. In the vector triangle, the ratio of the Fig. 4.15 Velocity diagram in the
vector to the sinc of the opposite angle cutting zone
will be equal. Mathematically,

vc
______________ vf vs
= ____ = __________
sin (90 – (f – a)) sin f sin (90 – a)

vc
_________ vf vs
fi = ____ = _____
cos (f – a) sin f cos a

sin f
vf = _________ Vc (4.3)
cos (f – a)

fi vf = vc × r
cos a
vs = _________ Vc (4.4)
cos (f – a)
194 Tool Design

Forces in metal cutting Metal cutting involves large amount of force. The power
supplied for cutting is utilised mainly in shearing the material and overcoming the
frictional forces between the tool and the chip. It is important to ascertain the cutting
forces acting during metal cutting, for calculation of power required in metal cutting,
in order to properly design cutting tools, for proper design of jigs and fixtures used to
hold the workpiece during the cutting operation, and for selection of suitable operating
conditions, like speed, feed and depth of cut.
A good understanding of the approach taken by metal-cutting researchers to the study
of cutting forces can be obtained by first limiting our studies to orthogonal cutting as
shown in Fig. 4.16a. Here the forces are all on one plane, and the components of the
forces can be added algebraically as in elementary mechanics.
Referring again to Fig. 4.16a, the resultant force R has two basic components Fc and
Ft. Fc, (cutting force) the force component in the direction of relative tool travel, deter-
mines the amount of work required to move the cutting tool to a given distance. The
component Ft (thoust force) does no work, but both components produce deflections
of the tool relative to the workpiece.
The force system of Fig. 4.16a is arrived at by assuming that the chip is a body in
stable mechanical equilibrium under the action of the forces exerted on it at the tool
face and at the shear plane. At the tool face, the force components F and N act on
the chip. F, known as the friction force, represents the frictional resistance met by the
chip as it slides over the face of the tool. N is known as the normal force. The ratio of
F to N is the coefficient of friction between chip and tool.
The force components acting at the shear plane are Fs and Fn. Fs represents the force
required to shear the metal on the plane of shear and is known as the shearing force.
Fn acts normal to the shear plane and results in a compressive stress being applied to
the plane of shear. The mean shear stress on the shear plane, which is equal to the
mean shear strength of the metal being cut, can of course be obtained by dividing Fs
by the area of the shear plane. Correspondingly, the compressive stress on the shear
plane is found by dividing Fn by the area of the shear plane.
The total work done by the cutting tool in removing metal, as determined by the value
of the force component Fc, is actually derived from two sources. It is the sum of the
work used in overcoming friction as the chip slides over the tool face and the work
consumed in shearing the metal on the shear plane.
The resultant force R acting on the tool can be resolved into two components Fc and
Ft. Fc and Ft can be determined by force dynamometers.
__› __› __›
R = Fc + Ft (4.5)
The rake angle, a can be measured from the tool, and forces F and N can then be
determined. The shear angle f can be obtained from its relation with chip thickness
ratio. Once f is determined, Fs and Fn can also be determined.
M. Eugene Merchant proposed a circle diagram of the forces known as Merchant’s
circle diagram (Fig. 4.16b), which is convenient to determine the relation between the
various forces and angles. In the diagram, two force triangles have been combined
and both the resultants of the frictional and shear force system together have been
replaced by R.
Design of Cutting Tools 195

Fig. 4.16 (a) Forces acting during metal cutting on chip and on the tool (b) Merchant’s
circle diagram of the metal cutting forces in the orthogonal cutting.

The frictional forces and the shear forces acting on the chip can be determined both
graphically and analytically using Merchant’s circle diagram. The procedure to con-
struct a Merchant’s circle diagram for graphical determination of the frictional and
shear forces is given below.
1. The cutting force (Fc) and the thrust (Ft) are measured using tool force dyna-
mometer. The cutting force (Fc) is drawn horizontally (along X-axis), and the
thrust force (Ft) is drawn vertically (along Y-axis). Joining of the two vectors
gives the resultant vector R.
2. With the centre of the resultant R, as centre and radius as half the resultant R,
a circle is drawn. This circle will enclose the endpoints of all the three vectors.
3. The rake angle is known from the geometry of the tool. A line is drawn extend-
ing over the rake surface of the tool with rake angle (a) from the vertical axis.
Extending this line will interest the circle at a point. Joining this point and the
resultant, R will give the friction vector (F).
196 Tool Design

4. A line can now be drawn from the head of the friction vector, to the head
of the resultant vector (R). This gives the normal vector (N). Therefore,
mathematically.
__› __› __› _› __›
R = Fc + Ft = F + N
5. The friction (b) is the angle between vector R and N.
6. The shear angle, f can be obtained from its relation with chip thickness ratio.
Once f is determined, Fs and Fn can also be determined. A line is drawn with
shear angle, f from the cutting force (Fc). This line intersect the circle at a point.
Joining this point and the resultant, R will give the shear Force (Fs).
7. Joining the endpoint of shear Force (Fs) and the other endpoint of the resultant
R will give the normal shear force (Fn). Therefore, mathematically,
__› __› __› __› __›
R = Fc + Ft = Fs + Fn
8. Scale and protractor can be used to measure the distances (forces) and
angles. D
Merchant’s circle diagram can also be
a
used to establish a relationship of the fric-
C
tional forces and the shear forces with the
cutting force and the thrust force. Figure (90 – a)
4.17 shows the frictional force system. f
E O
a Fc (b – a)
From Fig. 4.17, the relationship of the fric-
tion forces with the cutting force and the (90 – a) G
thrust force can be derived as follows: a
F = OA = CB = CG + GB = ED + GB
Ft R
fi F = Fc sin a + Ft cos a (4.6)
a F
N = AB = OD – CD = OD – GE
fi N = Fc cos a – Ft sin a (4.7)
b
Once the frictional forces are known the
coefficient of friction, m and the friction B N
angle b can be determined from the fol-
lowing relation: A
F Fig. 4.17 Frictional force system
m = tan b = __ (4.8)
N
Similarly from Fig. 4.18, the relationship of the shear forces with the cutting force and
the thrust force can be derived as follows:

Fn = AE = AD + DE = DE + CB
fi Fn = Fc sin f + Ft cos f (4.9)
Fs = OA = OB – AB = OB – BC
fi Fs = Fc cos f – Ft sin f (4.10)

Shear stress in metal cutting During metal cutting, the metal in severely com-
pressed in the area in front of the cutting tool. This causes high temperature shear
Design of Cutting Tools 197

a
A
Fs
(90 – f)
Fc f
C O
f (b – a)
(90 – f)
Fn D

E
Fig. 4.18(b) Shearing of chip in orthogonal
Fig. 4.18(a) Shear force system metal cutting

and plastic flow if the metal is ductile. When the stress in the workpiece just ahead of
the cutting tool reaches a value exceeding the ultimate strength of the metal, particles
will shear to form a chip element, which moves up along the face of the work. The
outward or shearing movement of each successive element is arrested by work hard-
ening and the movement transferred to the next element. The process is repetitive
and a continuous chip is formed.
The average shear stress, ts required during metal cutting can be mathematically
represented as,
Fs
ts = __ (4.11)
As
Where As is the cross-sectional area of the shear plane.
If A0 be the cross-sectional area of the chip before removal, then
A0 = w × t0 (4.12)
Where w is the width of cut
As is the cross-sectional area of the shear plane can also be expressed as,
A0
As = ____ (4.13)
sin f
From Eqs 4.11, 4.12 and 4.13,
Fs sin f
ts = _______ (4.14)
w × t0
Shear strain in metal cutting The expression of shear strain can derived using the
model of orthogonal cutting as shown in Fig. 4.18. The average shear strain, g can be
evaluated from the geometry,

AA¢ AE A¢E
g = ____ = ___ + ____
BE BE BE
198 Tool Design

= cot f + tan (f – a) (4.15)


This can be further simplified, and the expression of shear strain can also be expressed
as
cos a
g = _____________ (4.16)
sin f cos (f – a)
Work Done in Overcoming Friction Between Chip and Tool Wf The remainder of the
work done in removing a chip goes into overcoming the frictional resistance to sliding
of the chip on the tool.
Total Work Done in Cutting Wn Although the total amount of work done in cutting is
the sum of the work consumed in shearing the metal and that used in overcoming
friction.
The average tool designer, however, will seldom be concerned with orthogonal and
oblique cutting as viewed by the metal-cutting researcher. He is more interested in
the results of metal-cutting research that he can apply to his particular situation. The
following approach to metal-cutting forces is the result of metal-cutting research and
is presented in a manner that is usable by the average tool designer.
Fc
It has been stated earlier in this chapter that
most metal-cutting operations in industry are
comparable to oblique cutting; i.e., they are Ft
three-dimensional. The resultant force on Fr
the tool then has three basic components.
This situation exists in all cases except that
of orthogonal cutting. Figure 4.19 shows the
three basic components of the resultant force.
Note that the forces are established with ref- Fig. 4.19 References axis for cutting
erence to and along the basic machine axes. forces

In the lathe-turning operation shown in Fig. 4.19, there are three components.
Tangential force Fc This acts in the direction tangent to the revolving member and
is sometimes referred to as turning force. It is usually the highest of the three forces
and constitutes approximately 99 per cent of the total power required by the tool. The
tangential force is basically the cutting force. The power required in cutting PC = FC
× VC w.
Longitudinal/thrust force Ft This acts in a direction parallel to the axis of the work
and is sometimes referred to as the feeding force. It is reduced when high side-rake
angles are used. It averages about 40 per cent as high as the tangential force. The
longitudinal force is basically the thrust force or the feed force. The power require to
feed the tool Pt = Ft × longitudinal feed velocity.
Since the feeding velocity is very low, the power required is usually approximately one
per cent of the total.
Radial force Fr This acts in a radial direction from the centre of the work. It is the
force that holds the tool to the correct depth of cut. The radial force is the smallest of
the three tool forces, only 50 per cent as large as the longitudinal force. It requires no
power, in that there is no velocity in the radial direction. This force should be kept to
a minimum to reduce deflection, vibration, and chatter. This can be done by reducing
the nose radius and/or the side cutting-edge angle.
Design of Cutting Tools 199

These cutting forces of the tool point are usually measured with a dynamometer.
This device is one of the most reliable methods of measuring cutting forces because
mechanical efficiencies of the motor and machine are eliminated. It is inserted either
between the tool and the machine as a dynamometer-type toolholder or between
the work and the machine as a work-holding-type dynamometer. Dynamometer are
capable of measuring two or three force components at the same time, depending
upon the complexity of the dynamometer. The net power at the cutter and other prop-
erties can be determined from these force components.
The typical cutting-force dynamometer may be thought of as a spring scale. To illus-
trate the principle, a simple, one-component mechanical dynamometer is shown in
Fig. 4.20. This device is designed to measure the tangential cutting force by measur-
ing the amount of deflection of a cutting tool with a mechanical dial indicator. Such
a dynamometer would have to be precalibrated in order to determine force pounds
required to move the indicator needle.

0 T
15 5
10

Workpiece

Fig. 4.20 A simple one component mechanical dynamometer

The mechanical dynamometer shown in Fig. 4.20 serves to illustrate a principle. It


is not suitable for metal-cutting research as it must deflect a considerable amount to
function properly. The efficient toolholder must be rigid. Also the inertia of the moving
parts in the dial indicator does not give a steady reading.
Electric-resistance strain gauges are generally used as the measuring device in dyna-
mometers used in scientific metal-cutting research. Their use has made the design of
rigid and still sensitive recorders of cutting forces possible. The strain gauge consists
of a grid of fine wire bonded between two thin insulating strips of paper. Each end of
the grid is wired into a Wheatstone bridge circuit. The strain-gauge assembly is then
permanently cemented to either the cutting tool or a component of the dynamometer
that will be strained (deflected) by the forces involved in the cutting operation. Upon
being strained, the wire cross section of the strain gauge is changed. Its resistance
to change is in direct proportion to the strain or force being measured. With proper
calibration, the cutting forces in pounds can be read directly.
Power is the rate of doing work. The term power is defined as the work done divided
by the time during which this work is done, or
200 Tool Design

Fs
P = ___
t
where
F×s = work
F = force
s = distance moved
t = time

If the body on which the force is acting moves with velocity V, then
s
V = __
t
The total, P, required in metal machining is the summation of the power required in
cutting, the power required in longitudinal feed and the power required in radial feed,
(Pr if any):
P = Pc + Pt + Pr
Considering no radial feed, the above equation becomes
P = P c + Pt
fi P = Fc × Vc + Ft × longitudinal feed velocity
In comparison to the cutting velocity, the feed velocity is very nominal. Similarly, Ft
is very small compared to Fc. So the power required in feeding can be considered
negligible. Then the power required in metal machining is
P = F c × Vc

It is not usually practical or possible for the tool designer to calculate or measure the
forces involved in metal-cutting processes on the shop floor. The process engineer
or the shop foreman is usually more interested in the power required at the motor of
the machine tool than in the power required at the cutting tool. This is necessary to
determine the correct machine for the job. Without this knowledge the potential of a
machine tool may be wasted because a machine with a high power potential is used
to machine a workpiece that requires lesser peak loads. By the same token, a short-
age of power can be foreseen and overloading the machine prevented.
Perhaps a more practical approach for finding the power requirements in metal cut-
ting is by finding the specific energy of the material. Specific Energy, ut is defined as
the total energy per unit volume of material removed.

FcVc Fc Ws
ut = _____ = ___ ____3
wt0Vc wt0 mm

where w is the width of chip.


Therefore, specific energy is simply the cutting force to the projected area of cut.
The values of specific energy for different material is determined experimentally and
tabulated for practical use. Table 4.2 shows the specific energy for some common
materials. The specific energy is frequently used for estimation of the power require-
ments during metal cutting.
Design of Cutting Tools 201

Table 4.2 Average specific energy, ut (Ws/mm3)


Ferrous metals and alloys
Material Brinell hardness number
150-175 176-200 201-250 251-300 301-350 351-400

AISI:
1010-1025 1.58 1.83
1030-1055 1.58 1.83 2.18 2.62
1060-1095 – – 2.05 2.40 2.73
1112-1120 1.37
1314-1340 1.15 1.26 1.37
1330-1350 1.83 2.05 2.51 3
2015-2115 1.83
2315-2335 1.47 1.58 1.69 2.05 2.51
2340-2350 ... 1.37 1.58 1.91 2.27 2.73
2512-2515 1.37 1.58 1.83 2.18 2.51
3315-3130 1.37 1.58 1.91 2.27 2.73
3160-3450 ... 1.37 1.69 2.05 2.38 2.73
1430-4345 ... 1.26 1.58 1.91 2.27 2.73
4615-4820 1.26 1.37 1.58 1.91 2.27 2.38
5120-5150 1.26 1.37 1.69 2.05 2.38 2.73
52100 ... 1.58 1.83 2.27 2.73
6115-6140 1.26 1.47 1.83 2.27 2.73
6145-6195 ... 1.91 2.27 2.73 3.28 3.55
Plain cast iron 0.82 0.90 1.15 1.37
Alloy cast iron 0.82 1.15 1.47
Malleable iron 1.15
Cast steel 1.69 1.83 2.18

High-temperature alloys
Material Brinell hardness number ut (W-s/mm3)

A 286 165 2.24


A 286 285 2.54
Chromoloy 200 2.13
Chromoloy 310 3.22
Hastelloy-B 230 3
Inco 700 330 3.1
Inco 702 230 3
M-252 230 3
M-252 310 3.28
Ti-150A 340 1.77
U-500 375 3
4340 200 2.13
4340 340 2.54
202 Tool Design

Nonferrous metals and alloys


Matearial ut (W-S/mm3)

Brass:
Hard 2.27
Medium 1.37
Soft 0.90
Free-machining 0.68
Bronze:
Hard 2.27
Medium 1.37
Soft 0.90
Copper (pure) 2.46
Aluminium:
Cast 0.68
Hard (rolled) 0.90
Monel (rolled) 2.73
Zinc alloy (die-cast) 0.68

SOURCE: “Carboloy Application Data Manual,” Metallurgical Products Department, General Electric Company, by
permission.

It has been observed that machining with higher feed rates require less energy as
compared to machining with less feed rates. Therefore, in order to accurately use the
specific energy table, the specific energy values presented in the Table 4.2 need to be
multiplied by a feed correction factor. The feed correction factor values are presented
graphically in Fig. 4.21.
The power supplied by the motor of the metal cutting machine tool should be more
than the power required by the cutting tool for removing material, as a part of the
power supplied by the motor is used to overcome machine friction. Therefore, the
machine efficiency must be taken into account while calculating the power. The effi-
ciency value for machine tools is presented in Table 4.3.

Feed, mm/rev
0.025 0.05 0.01 0.127 0.259 0.500 0.01 1.53 2.54 3.81
2.0
Feed correction factor

1.6
1.5
1.0
0.8
0.6
0.5
0.4
0.3
0.001 0.002 0.004 0.005 0.010 0.020 0.040 0.060 0.100 0.150

Feed, in./rev.
Fig. 4.21 Feed correction factors based on normal tool geometry. When large lead angles
Design of Cutting Tools 203

and/or large nose radii are used, the chip will be thinned as a result of tool shape to
a chip thickness less than the actual feed set on the machine. Adjust as follows:

If tool cuts with a lead angle of Reduce the actual feed by this amount
and use the feed correction factor for
the finer feed
30° 10%
45° 30%
60° 50%
If the depth of cut is less than 30%
the nose radius

(From “Carboloy Application Data Manual,” Metallurgical Products Department,


General Electric Company)

Table 4.3 Typical overall machine-tool efficiencies


Values except for milling machines

Type Efficiency, %

Direct spindle drive 90


One-belt drive 85
Two-belt drive 70
Geared head 70

Milling machines

Rated power of machine (kW) Efficiency, %

2.23 40
3.73 48
5.59 52
7.46 52
11.19 52
14.92 60
18.65 65
22.38 70
28.84 75
37.30 80

SOURCE: ASTME, “Tool Engineers Handbook,” 2d ed., F W Wilson (ed.), McGraw-Hill, New York, 1959, by
permission.

During orthogonal machining of a mild steel part, a depth of


Example 4.1 cut of 0.8 mm is used at 55 rpm. If the chip thickness is 1.6 mm,
determine the chip thickness ratio. Also calculate the length of
chip removed in one minute if thework diameter is 50 mm before
the cut is taken. Assume a continuous type of chip.
Solution: Given: depth of cut, d = 0.8 mm, N = 55 rpm, chip thickness, tc = 1.6 mm,
work diameter D = 50 mm
204 Tool Design

Considering the depth of cut, d to be equal to the uncut chip thickness, t0, then

t0 = d = 0.8 mm

The chip thickness ratio, r


t0 0.8
r = __ = ___ = 0.5
tc 1.6

The length of chip to be removed in one minute, l0 will be

l0 = p DN = p × 50 × 55 mm/min = 8639.38 mm/min

Considering the volume of the material remains constant after deformation, then
volume of material before cut = volume of material after cut.

fi w × t0 × l0 = w × tc × lc (where w is the width of cut)

t0
fi lc = __ × l0 = r × l0
tc

= 8639.38 × 0.5 mm/min = 4319.693 mm/min.

In orthogonal turning of a mild steel bar of 60 mm diameter


Example 4.2 on a lathe, feed of 0.8 mm/rev was used. A continuous chip of
1.4 mm, thickness was removed at a rotational speed of 80 rpm.
Calculate the total length of chip removed in one minute.
Solution: Given: feed, f = 0.8 mm/rev, N = 80 rpm, chip thickness, tc = 1.4 mm, work
diameter D = 60 mm
For turning operation, feed is equal to the uncut thickness, t0, then

t0 = f = 0.8 mm
The chip thickness ratio, r
t0 0.8
r = __ = ___ = 0.57
tc 1.4
The length of chip to removed in one minute, l0 will be

l0 = p DN = p × 60 × 80 mm/min = 15079.64 mm/min


Considering the volume of the material remains constant after deformation, then vol-
ume of material before cut = volume of material after cut

w × t0 × l0 = w × tc × lc (where w is the width of cut)

t0
fi lc = __ × l0 = 15079.64 × 0.57 mm/min = 8595.397 mm/min.
tc
Design of Cutting Tools 205

Suppose you have the following data obtained from a metal


Example 4.3 cutting experiment (orthogonal machining). Compute the shear
angle, the shear stress, the coefficient of friction at the too/chip
interface, and the power consumed in metal cutting.
Machining data from 1020 steel, as received in air with a carbide tool, orthogonally
(tube cutting on lathe) with tube OD 73 mm. The cutting speed was 162 m/mm. The
tube wall thickness was 5 mm. The back rake was zero for all cuts.
Run Fc (N) Ft (N) Feed, mm/rev Chip thickness ratio, r

1 1468 1313 0.13 0.331


2 1371 1246 0.13 0.381
3 1824 1468 0.19 0.426
4 1869 1513 0.19 0.426
5 2269 1557 0.25 0.458
6 2491 1758 0.25 0.453

Solution: Given: rake angle, a = 0°, cutting speed, Vc = 162 m/mm, depth of cut/width
of cut, w = 5 mm
The shear angle for each experimental run (i.e., for each value of r), can be deter-
mined from the equation given below:
r cos a
f = tan–1 _________
1 – r sin a
In order to calculate the shear stress, the shear force, Fs for each experimental run
needs to be determined using the following equation:
Fs = Fc cos f – Ft sin f
For each value of the shear force Fs, shear stress, ts can be determined using the
following equation:
Fs sin f
ts = _______
w × t0
[Note: For orthogonal turning, feed is approximately equal to uncut chip thickness,
therefore, t0 = f ]
In order to calculate the coefficient of friction, the frictional force and normal force for
each experimental run need to be determined out using the following equations:
F = Fc sin a + Ft cos a
N = Fc cos a – Ft sin a
For each values of F and N, the coefficient of friction can be determined
F
m = tan b = __
N
The power consumed in cutting can be determined by the following equations:
Pc = Fc × Vc
The results for each experimental runs are tabulated as:
206 Tool Design

Run Fc Ft Feed Chip Shear Shear Frictional Normal Shear coeffi- P


(N) (N) mm/rev thickness angle Force, Force, Force, stress ts cient of (kW)
2
ratio, r f Fs F N (N/mm ) friction,
(N) (N) (N) m

1 1468.00 1313.00 0.13 0.33 18.31 981.05 1313.00 1468.00 474.28 0.89 237.82
2 1371.00 1246.00 0.13 0.38 20.86 837.54 1246.00 1371.00 458.76 0.91 222.10
3 1824.00 1468.00 0.19 0.43 23.07 1102.74 1468.00 1824.00 454.93 0.80 295.49
4 1869.00 1513.00 0.19 0.43 23.07 1126.50 1513.00 1869.00 464.74 0.81 302.78
5 2269.00 1557.00 0.25 0.46 24.61 1414.59 1557.00 2269.00 471.23 0.69 367.58
6 2491.00 1758 0.25 0.453 24.37 1543.63 1758.00 2491.00 509.57 0.71 403.54

In a machining operation that approximate orthogonal cutting,


Example 4.4 the cutting tool has a rake angle of 10°. The chip thickness
before the cut is 0.5 mm and chip thickness after the cut is 1.125
mm. Calculate the shear plane angle and the shear strain.
Solution: Given: rake angle, a = 10°, uncut chip thickness, t0 = 0.5 mm, chip thickness,
tc = 1.125 mm
The chip thickness ratio, r
t0 0.5
r = __ = _____ = 0.44
tc 1.125
We know the shear angle,
r cos a 0.44 cos 10
f = tan –1 _________ = tan–1 ____________ = 25.13°
1 – r sin a 1 – 0.44 sin 10
Shear strain, g is expressed as
g = cot f + tan (f – a) = cot 25.13 + tan (25.13 – 10) = 24.
In orthogonal turning of a 50 mm diameter mild steel bar on a
Example 4.5 lathe the following data were obtained: Raked angle = 15°, cut-
ting speed = 100 m/min, feed 0.2 mm/rev, cutting force = 1800 N.
Feed force = 600 N. Calculate the chip thickness ratio, shear
plane angle, and coefficient of friction, if the chip thickness is
0.3 mm.
Solution: Given: cutting speed, Vc = 100 m/min, rake angle, a = 15°, feed, f = 0.2 mm/
rev, cutting force, Fc = 1800 N, thrust force, Ft = 600 N, chip thickness, tc = 0.3 mm
For the case of orthogonal turning, feed is approximately equal to uncut chip thick-
ness, therefore,
t0 = f = 0.2 mm
The chip thickness ratio, r
t0 0.2
r = __ = ___ = 0.667
tc 0.3
We know the shear angle,
r cos a 0.667 cos 15
f = tan–1 _________ = tan–1 _____________ = 37.91°
1 – r sin a 1 – 0.667 sin 15
In order to calculate the coefficient of friction, the frictional force and normal force
need to be determined out using the following equations:
Design of Cutting Tools 207

F = Fc sin a + Ft cos a = 1800 × sin 15 + 600 × cos 15 = 1045.43 N


N = Fc cos a – Ft sin a = 1800 cos 15 – 600 sin 15 = 1583.38 N
For values of F and N, the coefficient of friction can be determined as,
F 1045.43
m = __ = _______ = 0.66.
N 1583.38
For orthogonal cutting of a component, the feed force was 750 N
Example 4.6 and cutting force was 1500 N. Find out the shear force and nor-
mal to shear force (compressive force) on the shear plane and
the coefficient of friction of the chip on the tool face. Assume
chip thickness ratio as 0.28 and tool rake angle as 10°.
Solution: Given: Chip thickness ratio, r = 0.28, rake angle, a = 10°, cutting force,
Fc = 1500 N, thrust force, F1 = 750 N
We know the shear angle,
r cos a 0.28 cos 10
f = tan–1 _________ = tan–1 ____________ = 16.16°
1 – r sin a 1 – 0.28 sin 10
We know, shear force, Fs can be expressed as,
Fs = Fc cos f – Ft cos f = 1500 × cos 16.16 – 750 × sin 16.16 = 1231.99 N
We know, normal to shear force (compressive force), Fn can be expressed as,
Fn = Fc sin f + Ft cos f = 1500 sin 16.16 + 750 cos 16.16 = 1137.85 N
In order to calculate the coefficient of friction, the frictional force, F and normal force,
N need to be determined out using the following equations:
F = Fc sin a + Ft cos a = 1500 × sin 10 + 750 × cos 10 = 999.08 N
N = Fc cos a – Ft sin a = 1500 cos 10 – 750 sin 10 = 1346.98 N
For values of F and N, the coefficient of friction can be determined as,
F 999.08
m = __ = _______ = 0.74.
N 1346.08
During orthogonal turning operation on a work piece of diam-
Example 4.7 eter 120 mm at 100 m/min with rake angle 15°, the width of the
cut and the chip thickness are 0.4 mm and 0.3 mm, respectively.
The feed during the operation was 0.2 mm/rev. If the cutting
force and the thrust force are 1200 N and 300 N, respectively,
calculate the shear angle, friction angle, shear stress and shear
strain.
Solution: Given: cutting speed, Vc = 100 m/min, rake angle, a = 15°, width of the cut,
w = 0.4 mm, chip thickness, tc = 0.3 mm, feed, f = 0.2 mm/rev, cutting force, Fc = 1200
N, thrust force, Ft = 300 N
For the case of orthogonal turning, feed is approximately equal to uncut chip thick-
ness, therefore,
t0 = f = 0.2 mm
208 Tool Design

The chip thickness ratio, r


t0 0.2
r = __ = ___ = 0.67
tc 0.3
We know the shear angle,

r cos a 0.67 cos 15


f = tan–1 _________ = tan–1 ____________ = 38.06°
1 – r sin a 1 – 0.67 sin 15
In order to calculate the coefficient of friction, the frictional force and normal force
need to be determined using the following equations:
F = Fc sin a + Ft cos a = 1200 × sin 15 + 300 × cos 15 = 600.36 N
N = Fc cos a – Ft sin a = 1200 cos 15 – 300 sin 15 = 1081.47 N
For values of F and N, the coefficient of friction can be determined as,
F 1045.43
m = __ = _______ = 0.555
N 1583.38
The friction angle can be determined using the following equations:
b = tan–1 m = tan–1 0.555 = 29.03°
Shear force, Fs is expressed as,
Fs = Fc cos f – Ft sin f = 1200 × cos 38.06 – 300 × sin 38.06 = 759.89 N
The average shear stress, ts required during metal cutting can be mathematically
represented as,
Fs sin f 759.89 sin 38.06
ts = _______ = ______________ N/mm2 = 5855.77 N/mm2
w × t0 0.4 × 0.2
Shear strain, g is expressed as
g = cot f + tan (f – a) = cot 38.06 + tan (38.06 – 15) = 1.703.
In an orthogonal cutting, the following data was recorded: chip
Example 4.8 length of 80 mm was obtained with an uncut chip length of 200
mm and the rake angle used was 20° and depth of cut 0.5 mm.
The horizontal and vertical component of forces was 2000 N
and 200 N, respectively. Determine the shear angle, chip thick-
ness and the friction angle.
Solution: Given: chip length, lc = 80 mm, uncut chip length, l0 = 200 mm, rake angle,
a = 20°, depth of the cut, d = 0.5 mm, cutting force, Fc = 2000 N, thrust force,
Ft = 200 N
Considering the depth of cut, d to equal to the uncut chip thickness, t0, then t0 = d =
0.5 mm
Considering the volume of the material remains constant after deformation, then vol-
ume of material before cut = volume of material after cut
w × t0 × l0 = w × tc × lc (where w is the width of cut)
t 0 lc 80
fi r = __ = __ = ____ = 0.4
tc l0 200
Design of Cutting Tools 209

Therefore, the chip thickness, tc will be


t0 0.5
tc = __ = ___ = 1.25 mm
r 0.4
The shear angle,
r cos a 0.4 cos 20
f = tan–1 _________ = tan–1 ___________ = 23.53°
1 – r sin a 1 – 0.4 sin 20
In order to calculate the coefficient of friction force, F and normal force, N need to be
determined using the following equations:
F = Fc sin a + Ft cos a = 2000 × sin 20 + 200 × cos 20 = 871.98 N
N = Fc cos a – Ft sin a = 2000 cos 20 – 200 sin 20 = 1810.98 N
For values of F and N, the coefficient of friction can be determined as,
F 871.98
m = __ = _______ = 0.48
N 1810.98
The friction angle can be determined using the following equation:
b = tan–1 m = tan–1 0.48 = 25.71°.
A mils steel tubing of 50 mm outside diameter is turned orthog-
Example 4.9 onally on a lathe with cutting speed of 20 m/min with a tool rake
angle of 35°. The tool is given a feed of 0.1 mm/rev and its is
found that the cutting force is 2500 N and feed force 1000 N.
Length of continuous chip in one revolutions is 80 mm. Calculate
the shear plane angle, coefficient of friction, chip thickness and
its velocity.
Solution: Given: Diameter, D = 50 mm, chip length, lc = 80 mm/rev, rake angle, a = 35°,
cutting speed, Vc = 20 m/mm, feed, f = 0.1 mm/rev cutting force, Fc = 2500 N, thrust
force, Ft = 1000 N
For the case of orthogonal turning, feed is approximately equal to uncut chip thick-
ness, therefore,
t0 = f = 0.1 mm
Lenth of an uncut chip, 10 for one revolution will be equal to the circumference of the
tube, therefore
l0 – t × D = 157.08 mm/rev
Chip thickness ratio,
t0 80
r = __ = lc/l0 = ______ = 0.51
tc 157.08
Therefore, the chip thickness, tc will be
t0 0.1
tc = __ = ____ = 0.196 mm
r 0.51
Therefore, the chip thickness is 0.196 mm
Chip flow velocity, Vr can be expressed as
210 Tool Design

m
Vf = r × Vc = 0.51 × 20 ____ = 102 m/min
min
The velocity of the chip is 102 m/min.
The shear angle,
r cos a 0.51cos 35
f = tan–1 ________ = tan–1 ____________ = 30.56°.
1– r sin f 1 – 0.51 sin 35

In order to calculate the coefficient of friction, the frictional force and normal force
need to be determined out using the following equations:
F = Fc sin a + Ft cos f = 2500 × sin 35 + 1000 × cos 35 = 2253.093 N
N = Fc cos a – Ft sin a = 2500 cos 35 – 1000 sin 35 1474.303 = N
For values of F and N, the coefficient of friction can be determined as,
F 2253.093
m = __ = ________ = 1.528.
N 1474.303
During orthogonal cutting of mild steel at 2 m/s with rake angle
Example 4.10 15°, the width of the cut and the depth of cut are 5 mm and 0.18
mm, respectively. The shear angle was measured to be 34°.
If the cutting force and the thrust force are 500 N and 200 N,
respectively, calculate the percentage of the total energy that
is dissipated in the shear plane during cutting.
Solution: Given: cutting speed, Vc = 2 m/s, rake angle, a = 15°, shear angle, f = 34°,
width of the cut, w = 5 mm, depth of cut, d = 0.18 mm, cutting force, Fc = 500 N, thrust
force, Ft = 200 N
The total power required in cutting (neglecting the power used in feeding the tool)
Ptotal, will be
Ptotal = Fc × Vc = 500 × 2W = 1000 W
We know, shear force Fs = Fc cos f – Ft sin f
Fs = 500 × cos 34 – 200 × sin 34 = 302.68 N
cos a cos 15
Shear velocity, Vs = _________ × Vc = ___________ × 2 = 2.04 m/s
cos (f – a) cos (34 – 15)
Energy consumed in shearing the material, Pshear = Fs × Vs = 302.68 × 2.04 W
= 618.43 W
The percentage of total energy dissipated in the shear plane
Pshear
_____ 618.43
× 100 = ______ × 100 = 61.84%.
Ptotal 1000

For turning hard aluminium alloy on a 15 KW lathe (mechanical


Example 4.11 efficiency 80%) at a width of cut of 6 mm, a rake angle of 10°,
and a cutting speed of 100 m/min, the coefficient of friction
was found to be 0.3. The chip thickness was 2 mm. What is your
estimate of the material’s shear strength?
Solution: Given: Ptotal = 15 KW, a = 10°, w = 6 mm, m = 0.3 and tc = 2 mm
Design of Cutting Tools 211

Net power transmitted, Pnet = 80% of Ptotal = 80% of 15 KW = 12 KW = 12 KW


= 12000 W
Considering the all the power transmitted is used in cutting (i.e. neglecting the power
used in feeding), then
100
Pnet = 1200 W ª Pc × Vc = Fc × ____ W
60
fi Fc = 7200 N
Specific energy, ut is defined as the total energy per unit volume of material removed.
Then,
FcVc Fc Ws
ut = _____ = ___ ____3
w t0Vc wt0 mm
FcVc Fc Nm
fi ut = _____ = ___ ____3
wt0Vc wt0 mm
From Table 4.2, assuming the specific energy value of hard aluminium to be
Ws
0.9 ____3 , then
mm
N Fc 7200
ut = 0.9 × 1000 ____2 = ___ = _____
mm wt0 6 × t0
fi t0 = 1.33 mm
The chip thickness ratio, r
t0 1.33
r = __ = ____ = 0.665
tc 2
r cos a
We know the shear angle, f = tan–1 _________
1 – r sin a
0.655 cos 10
f = tan–1 _____________ = 36.5°
1 – 0.655 sin 10
The friction angle can be determined using the following equations:
b = tan–1 m = tan–1 0.3 = 16.69°
From Merchant’s circle diagram,
Ft = Fc tan (b – a) = 7200 tan (16.9 – 10) = 871.30 N

Shear stress, Fs = Fc cos f – Ft sin f

Fs = 7200 cos 36.5 – 871.30 sin 36.5 = 5269.381 N


The average shear stress, ts required during metal cutting can be mathematically
represented as,
Fs sin f 5269.381 sin 36.5 N
ts = _______ = _______________ = 660.323 ____2
w × t0 6 × 1.33 mm
During metal cutting when the shear stress in the workpiece just ahead of the cutting
tool reaches a value exceeding the shear strength of the metal, material will shear to
form a chip element, which moves up along the face of the work. Therefore, the shear
strength of the material is equal to the shear stress, which is 660.323 N/mm2
212 Tool Design

In an orthogonal turning operation, the following cutting condi-


Example 4.12 tions apply: feed = 0.2 mm/rev, friction angle = 25°, rake angle =
15°, cutting speed = 1.5 m/s, chip thickness = 0.3 mm, depth of
cut = 2.5 mm, the shear strength of the material = 450 N/mm2.
Find the cutting force and thrust force.
Solution: Given: feed, f = 0.2 mm/rev, friction angle, b = 25°, rake angle, a = 15°, cut-
ting speed, Vc = 1.5 m/s, chip thickness, tc = 0.3 mm, depth of cut, d = 2.5 mm, shear
strength of the material, ts = 450 N/mm2.
For the case of orthogonal turning, feed is approximately equal to uncut chip thick-
ness, therefore,
t0 = f = 0.2 mm
The chip thickness ratio, r
t0 0.2
r = __ = ___ = 067
tc 0.3
r cos a
We know the shear angle, f = tan–1 _________
1 – r sin a
0.67 cos 15
f = tan–1 ____________ = 38.5°
1 – 0.67 sin 15
The average shear stress, ts required during metal cutting can be mathematically
represented as,
Fs sin f
ts = _______
w × t0
ts × w × t0 450 × 2.5 × 0.2
fi Fs = _________ = _____________ = 365 N
sin f sin 38.5
From Merchant’s circle diagram,

Fs = Fc sec (b – a) cos (f + b – a)

From the above equation Fc, can be expressed in terms of Fs, as


cos (b – a)
Fc = ____________ × Fs
cos (f + b – a)

cos (25 – 15)


= _________________ × 365
cos (38.5 + 25 – 10)
cos 10 × 365
= ___________ = 542.48 N
cos 48.5
We know, the shear force
Fs = Fc cos f – Ft sin f

From the above equations Ft, can be expressed in terms of Fs, as


Fc cos f – Fs 542.48 cos 38.5 – 365
Ft = ___________ = __________________ = 95.65 N.
sin f sin 38.5
Design of Cutting Tools 213

A 100 cm long, 40 mm diameter 304 stainless steel rod is going


Example 4.13 to be machined at the lathe into the geometry shown below
(Fig. E4.1). The total given is of HSS material with a rake angle
of 10°. The friction coefficient between the total and the mate-
rial given is 0.5. The shear angle, during the metal cutting was
found to be 36.9°. Find out the turning parameters (Depth of
cut, feed cutting speed, MRR). Rotational speed, N = 400 rpm,
Tool travel speed along the workpiece length, 30 mm/min, and
the machining is going to be done by one pass. If the specific
energy of the steel given as 4.1 W.s/mm3, what is the cutting
force? Also calculate the chip thickness and its speed?

Fig. E4.1 The finished steel rod

Solution: Given: Initial diameter, Di = 40 mm, Final diameter, D0 = 35 mm, coefficient


of friction, m = 0.5, rake angle, a = 10°, shear angle, f = 36.9°, Rotational speed,
N = 400 rpm, feed rate, f × N = 30 mm/min, Specific Energy, ut = 4.1 W.s/mm3
Let us assume that the turning to be orthogonal metal cutting
Feed,
feed rate 30
f = ________ = ____ = 0.075 mm/rev
N 400
For orthogonal turning, feed is approximately equal to uncut chip thickness, therefore,
t0 = f = 0.075 mm
Original diameter of the workpiece, Di = 40 mm
Diameter of the workpiece after machining, Do = 35 mm
As the machining is done in one pass, so from the figure given,
Depth of cut, d = (Di – Do)/2 = 2.5 mm
Average diameter, Davg = 37.5 mm
Cutting speed, Vc can be expressed as
p DavgN p × 37.5 × 400
Vc = _______ = ____________ m/min = 47.123 m/min
1000 1000
Material Removal Rate, MRR for turning can be expressed as
47.123 × 103 mm3
MRR = Vc × f × d = ___________ × 0.075 × 2.5 ____ = 147.26 mm3/s
60 s
We know power required in cutting, Pc is given by
Pc = MRR × ut = 147.26 × 4.1 W = 603.766 W
214 Tool Design

Also,
P c = F c × Vc
Pc
fi Fc = __ = 603.766 × 60/47.123 N = 768.75 N
Vc
Therefore, the cutting force is 768.75 N
We know chip thickness ratio, r can be expressed as
t0 sin f sin 36.9
r = __ = _________ = _____________ = 0.673
tc cos (f – a) cos (36.9 – 10)
0.075
tc = _____ = 0.111 mm
0.673
Therefore, the chip thickness is 0.111 mm
Chip flow velocity, Vf can be expressed as
m
Vf = r × Vc = 0.673 × 47.123 ____ = 31.713 m/min
min
The velocity of the chip is 31.173 m/min.
Find the power required at the motor for turning operation with
Example 4.14 the following conditions:

Machine: Geared head turret lathe


Workpiece material: AISI 1020 steel, BHN 175
Cutting Speed: 91 m/min
Feed: 0.65 mm/rev
Depth of cut: 6.5 mm
Cutting tool: Carbide steel-cutting grade with 30° side cutting-edge angle
Solution: Given: cutting speed, Vc = 91 m/min, feed, f = 0.65 mm/rev, and depth of cut/
width of cut, w = 6.5 mm
For workpiece material: AISI 1020 steel, BHN 175, the specific energy can be obtained
from Table 4.2
Ws
ut = 1.58 ____2
mm
For 30° side cutting-edge angle, the actual feed for finding the feed correction factor
will be reduced by 10% (refer to Fig. 4.21)
For corrected feed of 0.585 mm/rev, the feed correction factor is about 0.9 (refer to
Fig. 4.21)
Therefore, the corrected specific energy will be,
ut = 1.58 × 0.9 = 1.42 Ws/mm3
We know power required in cutting, Pc is given by
Pc = MRR × ut = Vc × f × d × ut
Design of Cutting Tools 215

91 × 1000
= _________ × 0.65 × 6.5 × 1.42 = 9099.24 W
60
From Table 4.3, the efficiency of a all geared lathe is 70%, then
The power required at the motor, P will be
Pc 9099.24
P = ____ = _______ = 12998.91 W.
0.70 0.70
During orthogonal turning of mild steel at 210 m/min with rake
Example 4.15 angle 12°, the uncut chip thickness and feed are 0.8 mm and
0.45 mm/rev solid respectively. The width of cut was 2 mm. If
the average value of the coefficient of friction between the tool
and the chip is 0.4 and the shear strength of the work material
is 39x 107 N,m2, calculate the cutting and thrust components of
the machining force.
Solution: Given: rake angle, a = 12°, cutting speed, Vc = 210 m/min, chip thickness,
tc = 0.8 mm, feed, f = 0.45 mm/rev, width of cut, w = 2 mm, coefficient of friction,
m = 0.4, shear strength of the material, ts = 390 N/mm2.
For the case of orthogonal turning, feed is approximately equal to uncut chip thick-
ness, therefore,
t0 = f = 0.45 mm
The chip thickness ratio, r
t0 0.45
r = __ = ____ = 0.56
tc 0.8
r cos a
We know the shear angle, f = tan–1 _________
1 – r sin a
0.56 cos 12
f = tan–1 ____________ = 31.796°
1 – 0.56 sin 12
The average shear stress, ts required during metal cutting can be mathematically
represented as,
Fs sin f
ts = _______
w × t0
ts × w × t0 390 × 2 × 0.45
fi Fs = _________ = ____________ = 666.165 N
sin f sin 31.796
The friction angle can be determined using the following equations:
b = tan–1 m = tan–1 0.4 = 21.8°
From Merchant’s circle diagram,
Fs = Fc sec (b – a) cos (f + b – a)
From the above equation Fc, can be expressed in terms of Fs as
cos (b – a)
Fc = ____________ × Fs
cos (f + b – a)
216 Tool Design

cos (21.8 – 12)


= ____________________ × 666.165
cos (31.796 + 21.8 – 12)

cos 9.8
fi Fc = ________ × 666.165 = 877.70 N
cos 41.59

Shear force, Fs can be expressed as,

Fs = Fc cos f – Ft sin f

From the above equation Ft, can be expressed in terms of Fs as

Fc cos f – Fs
Ft = ___________
sin f

877.70 × cos 31.796 – 666.165


= __________________________ = 151.487 N.
sin 31.796

Factors affecting forces and power The amount of power used in a machining oper-
ation is generally proportional to the cutting speed, in that the rate of metal removal
is proportional to the speed. Speed, however, has little effect on cutting forces within
the normal cutting range. Notably higher forces may result at extremely slow speeds,
but as the speed approaches the normal cutting range, the forces begin to level
off, as shown in Fig. 4.22. Thereafter, there is a slight decrease in forces as the

Fig. 4.22 Cutting speed vs. force (Metallurgical Product Department,


General Electric Company)
Design of Cutting Tools 217

cutting speed is increased. This is probably due to an increase in shear angle, which
results in a shorter shear plane and a reduction in chip thickness. For all practical pur-
poses, the change in cutting forces due to increased speed may be ignored as long
as the speed is within the normal cutting range.
Feed causes an increase in all three forces, as shown in Fig. 4.23. The depth of cut
increases tangential and longitudinal forces but makes no changes in the radial force
when using a 0° lead-angle too (see Fig. 4.24). The radial force is increased when
tools with increasingly larger side cutting edge angles are used.

Fig. 4.23 Effect of feed on the components of the tool force when turning 0.21 carbon steel
with a high-speed-steel tool. (Tool designation, 8, 14, 6, 6, 6, 0, 3.2. Depth of cut,
3.2 mm; cutting speed, 24 m/min) (Metallurgical Product Department, General
Electric Company)

Comparison of Fig. 4.23 and 4.24 shows that the depth of cut has the most direct
effect on cutting forces. Therefore, it is possible to remove metal more efficiently at
heavy feeds since less power will be required.
Figure 4.25 shows the effect of rake angles on cutting forces. Back rake has the least
effect on forces, while positive side rake has the greatest effect. Increased side-rake
angles produce greater reduction of forces when cutting at low speeds than at high
speeds. When comparing b and c, it can be seen that an increase in speed has a
definite influence on reducing the longitudinal tool force. For this reason, tools with
negative rakes should be operated at higher cutting speeds.
218 Tool Design

Figure 4.27 shows that increased nose radii result in a decided increase in radial force
and some increase in tangential force. Larger radii prolong too life but cause greater
tool forces through less efficient cutting and tendency to chatter.

Fig. 4.24 Effect of depth of cut on tool-force components. (Test conditions same as
Fig. 4.23 except that a constant feed of 0.4 min/rev. was used) (Metallurgical
Product Department, General Electric Company)
Design of Cutting Tools 219

Fig. 4.25 The influence of rake angles on cutting forces (Metallurgical Product Department,
General Electric Company)

An increase in the side cutting-edge angle increases the radial force, as shown in
Fig. 4.26. A change from 0 to 45° doubles the radial force. Efficiency of metal cutting is
slightly reduced because increased side cutting-edge angles cause wider but thinner
chips to be formed through less efficient cutting and tendency to chatter.

Fig. 4.26 Effect of side cutting-edge angle on cutting forces (Metallurgical Product
Department, General Electric Company)

One of the most important factors in determining tool forces and power is the work
material itself. Ductility, hardness, coefficient of friction, and work hardening all have
definite effects on tool forces. Increased ductility results in increased chip thickness
and therefore increased forces. Higher hardness causes higher power or force, as
shown in Fig. 4.25 and 4.26. The coefficient of friction can be reduced by additions
such as sulfur and lead, and lower forces will result. Austenitic material work hardens
and requires higher forces and power.
220 Tool Design

Fig. 4.27 Effect of nose radius on cutting forces. Note that radial force increases form 50
lb with 0° nose radius to 175 lb with a 1/4-in. radius (Metallurgical Product
Department, General Electric Company)

Fig. 4.28 Specific energy, W-s/mm3 when turning various metals. (Tool designation, 8, 14,
6, 6, 6, 0, 3, except as indicated for aluminium; depth of cut, 2.54 mm; feed, 0.32
mm/rev) (Metallurgical Product Department, General Electric Company.)
Design of Cutting Tools 221

Fig. 4.29 Specific energy when turning steel. (Metallurgical Product


Department, General Electric Company)

4.5 GENERAL CONSIDERATION FOR


METAL CUTTING
Metal is machined in order to remove surplus metal and bring the workpiece to the
desired size. This appears to be a simple statement, but in reality the metal-removal
process is quite complex. Many variables enter in and cause a cutting tool to react
in a certain way. Examples are the workpiece material, the cutting-tool material, the
rigidity of the machine, the rigidity of the workpiece, the rigidity of the setup, the cut-
ting feed and speed, chatter, tool wear, and chip control. All these and more have an
effect on the metal-cutting operation. It is the purpose of this section to discuss these
factors so that the tool designer can have a better understanding of the metal-cutting
process.

Machinability of various materials The preceding discussion of the effect of work


material on cutting forces and power leads into the general discussion of the machin-
ability of various materials. The term machinability is used when describing the rela-
tive ease of machining a metal. It is usually expressed as a numerical machinability
rating, although it has been expressed as relative cutting speeds (see Tables 4.4 and
4.5). Machinability ratings are determined by comparing the relative machinability of
a material with another common easily machined material as an index. Steels are
rated by comparison with cold drawn B1112. The B1112 grade is rated at 100 per
cent machinability “when turned with a suitable cutting field 55 m/min under normal
cutting conditions.”
Machinability ratings are based upon the relative speeds used with each given mate-
rial to obtain a given tool life. They are to be used as a guide in selecting the first
cutting speeds when machining a new material or as an aid in the selection of work-
piece material. In theory, a material whose rating is 50 should be machined at half
Table 4.4 Machinability rating of various metals

Class I: ferrous (70% and higher) Class II: ferrous (55-65%) Class III: ferrous (40-50%) Class IV: ferrous (40% or lower)

AISI Rating, % BHN AISI Rating, % BHN AISI Rating, % BHN AISI Rating, % BHN
222 Tool Design

C1109 85 137-166 C1141 65 183–241 C1008 50 126–163 A2515† 30 179-229


C1115 85 143-179 C1020 65 137–174 C1010 50 131–170 E3310† 40 170-229
C1117 85 143-179 C1030 65 170–212 C1015 50 131-170 E25100‡ 30 183-229
C1118 80 143-179 C1035 65 174-217 C1050† 50 179-229 E9315 40 179-229
C1120 80 143-179 C1040† 60 179-229 C1070† 45 183-241 Ni resist† 30
C1132 75 187-229 C1045† 60 179-229 A1320 50 170-229 Stainless 18*8† 25 150-160
C1137 70 187-229 A2317 55 174-217 A1330 50 179-235 (austenitic)
C1022 70 159-192 A3120 60 163-207 A1335† 50 187-241 Manganese oil-
C1016 70 137-174 A3130† 55 179-217 A1340† 45 187-241 hardening steel‡ 30
B1111 95 179-229 A3140† 55 187-229 A2330† 50 179-229 Tool-steel, low
B1112 100 179-229 A4032† 65 170-229 A2340† 45 187- 241 tungsten, chrome,
B1113 135 179-229 A4037† 65 179-229 A3145† 50 187-235 and carbon† 30 200-218
A4023 70 156-207 A4042† 60 183-235 A4150† 50 187-235 High-speed steel‡ 30
A4027* 70 166-212 A4047† 55 183-235 A4340† 45 187-241 High-carbon-high-
Malleable: A4130† 65 187-229 A6120 50 179-217 chrome tool steel‡ 25
Standard 120 110-145 A4137† 60 187-229 A6145† 50 179-235
Pearlitic 90 180-200 A4145† 55 187-229 A6152† 45 183-241
Pearlitic 80 200-240 A4320 55 179-228 A9260† 45 187-255
Cast iron, soft 80 160-193 A4615 65 174-217 NE9261† 50 179-217
Cast steel (0.35 C) 70 170-212 A4640† 55 187-235 Ingot iron 50 101-131

Contd...
Contd...

Class I: ferrous (70% and higher) Class II: ferrous (55-65%) Class III: ferrous (40-50%) Class IV: ferrous (40% or lower)

AISI Rating, % BHN AISI Rating, % BHN AISI Rating, % BHN AISI Rating, % BHN

Stainless iron (12% A4815 50 187-229 Wrought iron 50 101-131


Cr, free-machining) 70 163-207 A5120 65 170-212 Stainless (18-8 free-
A5140† 60 174-229 machining) 45 179-212
A5150† 55 179-235
A8620 or 8720 60 170-217
A8630† or 8730† 65 179-229
A8740† or 8640† 55 183-235
A8745† or 8645† 50 183-241
A9140† 60 179-229
Cast iron, hard 50 220-240
Medium 65 193-220

*For Independent Research Committed by E Slaughter, E J Hergenrcether, and O W Boston. Each metal is rated to the nearest 5% based on the4 100% rating of AISI steel specification B1112,
cold-rolled or cold-drawn, when machined with a suitable cutting fluid at 55m/mm unser normal cutting conditions using high-speed tools.

NOTE: The terms “annealed” (†) and “spheroidised anneal” (‡) refer specifically to the commercial practise in steel mills, prior to cold drawing or cold rolling, in the production of the steels
specifically mentioned.

SOURCE: ASTME, “Tool Engineers Handbook,” 2d ed., F W Wilson (ed.), McGraw-Hill, New York, 1959, by permission.
Design of Cutting Tools 223
224 Tool Design

Table 4.5 Machinability rating by cutting speeds for steel and cast iron
Alloya Condition or Structureb BHN Speed no.c
b
heat treatment Steels HSSd Carbidee
1112 Hot-rolled 10% P, 90% F 135 67 244
1020 Annealed 10% P, 90% F 115 46 168
8620 Annealed 30% P, 70% F 135 46 146
3140 Annealed 75% P, 30% F 190 23 85
3140 Q and T Temp mart 302 17 61
4140 Annealed 65% P, 35% F 180 24 101
4140 Annealed 90% P, 10% F 192 23 69
4140 Spher Spher 166 38 171
4140 Q and T Temp mart 300 17 64
4140 Q and T Temp mart 400 11 58
4340 Annealed 100% P 221 21 79
4340 Spher Spher 206 24 114
4340 Q and T Temp mart 300 19 79
4340 Q and T Temp mart 400 9 69
4340 Q and T Temp mart 500 5 38
5140 Annealed 80% P, 20% F 192 26 73
8640 Annealed 50% P, 50% F 170 38 99
8640 Annealed 75% P, 25% F 190 27 85
8640 Spher Spher 180 40 114
8640 Hot-rolled W 250 26 95
8640 Q and T Temp mart 300 18
8640 Q and T Temp mart 400 12 46
52100 Spher Spher 190 31 84
1345 Annealed 80% P, 20% F 207 24 88
4817-H Annealed W spher 217 27 137
9262-H Annealed 100% P 255 18 46
14B17 Annealed 60% P, 40% F 167 46 90
94B17 Annealed 40% P, 60% F 197 46 165
80B40 Annealed 100% P 212 24 79
81B40 Annealed 95% P, 5% F 207 24 88
86B45 Annealed 95% P, 5% F 212 20
98B40 Annealed 100% Pg 202 24 101
50B60 Annealed 100%h 205 29 67
Cast 1020 Annealed 30% P, 70% F 122 49 122
Cast 1020 Normalised 30% P, 70% F 134 41 70
Cast 1040 Dbl norm 70% P, 30% F 185 40 122
Cast 1040 Norm and ann 60% P, 40% F 175 41 116
Cast 1040 Normalised 70% P, 30% F 190 37 99
Design of Cutting Tools 225

Alloya Condition or Structureb BHN Speed no.c


heat treatmentb Steels HSSd Carbidee

Cast 1040 Norm and O Q 80% P, 20% F 225 24 95


Cast 1330 Normalised 60% P, 40% F 187 23 43
Cast 1330 Norm and temp 60% P, 40% F 160 37 70
Cast 4130 Annealed 50% P, 50% F 175 29 79
Cast 4130 Norm and spher spher 175 27 61
Cast 4340 Norm and ann 85% P, 15% F 200 18 64
Cast 4340 Norm and spher spher 210 29 88
Cast 4340 Q and T Temp mart 300 14 61
Cast 4340 Q and T Temp mart 400 11 55
Cast 8430 N and T 1200°F Fine spher 200 27 61
Cast 8430 N and T 1275°F Fine spher 180 34 73
Cast 8630 Normalised 70% P, 30% F 240 23 55
Cast 8030 Annealed 55% P, 45% F 175 37 88
Cast irons
Grey iron Annealed 100% F 100 – 224
Grey iron As-C slow cool Coarse P 195 – 98
Grey iron As-C fast cool Fine P 225 – 82
Grey ironi As-C Acicular 262 – 52
Nodular 1 As-C 80% P, 20% F 265 – 67
Nodular 2 As-C 40% P, 60% F 215 – 98
Nodular 3 As-C 40% P, 60% F 207 – 116
Nodular 1 Annealed 3% P, 97% F 183 – 152
Nodular 3 Annealed 100% F 170 – 238
a
Carburising steels (0.20%C) are usually received in the hot-rolled condition; medium-carbon steels, either hot-rolled
or annealed.
b
Abbreviations used in this table: Ann. annealed; As-C, as-cast; Dbl, double; F, ferrite; mart, martensite; N or norm, nor-
malised; O, oil; P, pearlite; Q, quenched, Spher, spheroidizd or spheroidite; T or Temp, tempered; W, Widmanstatten.
c
Specific-speed is the cutting speed in m/min which causes a given flank-wear land on the tool in 60 min. Flank-wear
land on high-speed-steel tools was 15 mm on carbide tools, 0.4 mm for cutting steel and 0.08 in. for cutting cast iron. All
the results are for turning operations. The depth of cut was 1.6 mm and feed 0.23 mm/rev for high-speed-steel tools; for
carbide tools, depth of cut was 2.5 mm and feed was 0.25 mm/rev.
d
18-4-1 high-speed steel.
e
General purpose cast-iron grade of carbide for cutting cast irons; general-purpose steel grade for steels.
f
Partly spheroidised. g Coarse pearlite. h Pearlite spheroidised.
i
High-alloy cast iron
SOURCE: ASTME, “Tool Engineers Handbook,” 2d ed., F W Wilson (ed.), McGraw-Hill, New York, 1959, by
permission.
226 Tool Design

the cutting speed used for the material whose rating is 100, assuming all factors are
equal. However, it is a well-known fact that production conditions are far from being
equal to conditions in the laboratory, where machinability ratings are determined. The
workpiece and tool rigidity may not be the same, the production machines may be
worn and loose, or the tool geometry may be different. It must be remembered that
machinability data are relative and therefore must be used as a relative guide. Usually,
higher the machinability rating, longer the tool life, less power required, and better the
surface finish. Machinability ratings may be of great value when applied in this manner
and relying on judgment and past experience.
Specific cutting speed is defined as the cutting speed corresponding to the predeter-
mined tool life.
Machinability rating, MR is the ratio of the specific cutting speed for certain tool life of
the test material to the specific cutting speed for same tool life of a standard material.
It is expressed as percentage. Mathematically,
V tmin
MR = ____ × 100% (4.17)
V smin
Machinability ratings can be very easily understood with the help of an example, For
example, if the tool life during a turning operation under standard condition (of feed,
depth of cut, tool material and tool geometry) is found to be 60 min at a cutting speed
of 100 m/min, the specific cutting speed for 60 min tool life, V60 = 100 m/min. Further,
if V t60 is the specific cutting speed for 60 min tool life for a test material and V t60 is the
corresponding specific speed for a standard material, then the machinability rating,
MR of the test material is given by
t
V60
MR = ___ × 100 %
Vs60
The reference material for steels, AISI 1112 steel has an index
Example 4.16 of 1. Machining of this steel at cutting speed of 0.5 m/s gives
tool life of 60 min. For the austenitic 302 SS steel, machining
at cutting speed of 0.23 m/s gives tool life of 60 min. And for
a tool life of 60 min, the AISI 1045 steel should be machined at
0.36 m/s. Find out the machinability of these steel in terms of
their machinability ratings.
Solution: Given: V s60 = 0.5 m/s.
t
For the austenitic 302 SS steel, V60 = 0.23 m/s,
Therefore, the machinability rating,
0.23
MR = ____ × 100% = 46%
0.5
t
For AISI 1045 steel, V60 = 0.36 m/s,
Therefore, the machinability rating,
0.36
MR = ____ × 100% = 72%
0.5
This index is smaller than 100%, therefore, AISI 1045 steel has a worse workability
than AISI 1112, but better than 302 SS.
Design of Cutting Tools 227

So, we can rate these steels in a descending order of macbinability.


AISI 1112 > AISI 1045 > 302 SS.
A series of tool life tests are conducted on two work materials
Example 4.17 under identical conditions, varying only speed in the test pro-
cedure. The first material, defined as the base material, yields
a Taylor’s life equation as VT 0.28 = 350, and the other material
(test material) yields a Taylor’s life equation as VT 0.27 = 440,
where speed is in m/min and the tool life in min. Determine the
machinability rating of the test material using cutting speed
that provides a 60 min too life as the basis of comparison.
Solution: For reference material:
V s60 T 0.28 = 350

350 _____
350
fi V s60 = ____ = = 111.22 m/min
T0.28 600.28
For test material:
440 _____
440
V t60 = ____ = = 145.66 m/min
T 0.27 600.27
Then the machinability rating, MR of the test material is given by
t
V60 145.66
MR = ___
s
× 100% = ______ × 100% = 131%
V60 111.22
Hence, the required machinability rating for the test material for 60 min tool life is
131%.
A machinability rating to be determined for a new work mate-
Example 4.18 rial using the cutting speed for a 60 min tool life as the basis
of comparison. For the base material (B1112 steel), test data
resulted in Taylor’s tool life equation parameter values of
n = 0.29 and C = 500, where speed is in m/min and tool life
in min. For the new material, the parameter values were n =
0.21 and C = 400. These results were obtained using cemented
carbide tooling. (a) Compute a machinability rating for the new
material. (b) Suppose the machinability criterion was the cut-
ting speed for 10min tool life. Compute the machinability rating
for this case. (c) What do the results of the two calculations
show about the difficulties in machinability measurements?
Solution: Given that the base or standard material is B1112 steel with the parameter
values for the Taylor’s tool life equation as n = 0.29 and C = 500
The new material whose machinability rating is to be determined has the following
parameter values for the same equation where n = 0.21 and C = 400.
The Taylor’s tool life equation is given by
VTn = C
Where,
V = cutting velocity employed, m/min
228 Tool Design

T = Tool life in minutes


C = constant; the cutting velocity for 1 min of elapsed time before reaching the
wear limit of the tool
n = constant which is considered a characteristics of the tool material, called
tool life index.
Note: for T = 1 minute, C becomes equal to the cutting speed
Each combination for work piece, tool material and cutting condition has its own n and
C values, both of which are determined experimentally
(a) The basis of comparison is 60 min tool life.
Thus, we have for the standard material

V s60 × T n = C
C 500
fi V60s
= ___n = _____ = 152.153 m/min
T 600.29
Similarly, for the test material, using the similar nomenclature and 60 min. Tool life,
we have
C 400
t
V60 = __n = _____ = 169.296 m/min
T 600.21
Then the machinability rating, MR of the test material is given by
t
V60 169.296
MR = ___
s
× 100% = _______ × 100% = 111%
V60 152.513
Hence, the required machinability rating for the new material for 60 min tool life is
111%.
(b) The machinability criterion has been changed to 10 min. Tool life. Thus, we have
The basis of comparison is 10 min tool life.
Thus, we have for the standard material
s
V10 × Tn = C
C 500
fi V10s
= ___n = _____ = 256.43 m/min
T 100.29
Similarly, for the test material, using the similar nomenclature and 10 min. Tool life,
we have
C 400
V t10 = ___n = _____ = 246.638 m/min
T 100.21
Then the machinability rating, MR of the test material is given by
t
V10 246.638
MR = ___
s
× 100% = _______ × 100% = 96.18%
V10 256.43
Hence, the required machinability rating for the new material for 10 min tool life is
96.18%.
(c) Comparing the results obtained in part a) and part b), it is seen that there is a
major drop in the value of the machinability rating of the test material or the new
Design of Cutting Tools 229

material as the tool life is decreased from 60 min to 10 min. Thus, it is thus obvi-
ous the machinability will vary for different specific speeds, even for same materials.
Therefore, in order to obtain meaningful comparison standard material is used for
reference (normally B1112 grade steel is taken) with specific speeds for 60 min tool
life for both the test and reference material.
A machinability rating is to be determined for a new work mate-
Example 4.19 rial using the cutting speed for a 60 min tool life as the basis
of comparison. For the base material (B1112 steel), test data
resulted in Taylor’s equation parameter values of n = 0.27 and
C = 450, where speed is in m/min and tool life is min. For the
new material and parameter values were n = 0.22 and C = 420.
These results were obtained using cemented carbide tooling.
(a) Compute a machinability rating for the new material using
cutting speed for a 30 min tool life as the basis of compari-
son. (b) If the machinability criterion were a tool life for a cut-
ting speed of 150 m/min, what is the machinability for the new
material?
Solution: Given that the base or standard material is B1112 steel with the parameter
values for the Taylor’s tool life equation as n = 0.27 and C = 450
The new material, whose machinability rating is to be determined has the following
parameter values for the same equation where n = 0.22 and C = 420.
(a) The basis of comparison is 30 min tool life.
Thus, we have for the standard material

V s30 × Tn = C
C 450
fi V s30 = ___n = _____ = 179.636 m/min
T 300.27
Similarly, for the test material, using the similar nomenclature and 30 min. Tool life,
we have
C 420
t
V30 = ___n = _____ = 198.739 m/min
T 300.22
Then the machinability rating, MR of the test material is given by
t
V30 198.739
MR = ___
s
× 100% = _______ × 100% = 110.63%
V30 179.636
Hence, the required machinability rating for the new material for 30 min tool life is
110.63%
(b) The machinability criterion has been changed to the tool life for a cutting speed of
150 m/min. Thus, the tool life for a cutting speed of 150 m/min

VTs × T n = C
C
fi T n = ___s
VT
450
fi T 0.27 = ____
150
230 Tool Design

1
____
fi T = 3 0.27 = 58.49 min

The basis of comparison is 58.49 min tool life.


For the test material, specific speed for 58.49 min tool life, we have
C 420
t
V58.49 = ___n = ________ = 171.589 m/min
T 58.490.22
Then the machinability rating, MR of the test material is given by
t
V58.49 171.589
MR = _____
s
× 100% = _______ × 100% = 114.39%
V58.49 150

Hence the required machinability rating for the new material for 58.49 min tool life is
114.39%
Many production people talk about machinability without really understanding the fac-
tors affecting it. To explain all the many factors requires the knowledge of an expert
metallurgist and is therefore beyond the scope of this discussion. However, a basic
understanding can be imparted by speaking in general terms and relating the discus-
sion to everyday observations of metal cutting.
It is common knowledge that soft and gummy materials do not machine as readily as
those harder and less gummy. This observation is verified from the information given
in Table 4.6 under the machinability ratings of aluminium-base alloys. Note that the
machinability rating is higher for the harder alloys. This is because soft aluminium
tends to adhere to the face of the cutting tool. The chip material begins to pile up on
the tool, resulting in a thicker chip and greater length of the shear plane. The surface
finish is poor because of the built-up edge on the cutting tool.

Table 4.6 Machinability rating of various metals and alloys


AISI no. or commo- BHN Machinabi- Aluminium Association BHN Machinabi-
nly used trade name lity rating* designation lity rating*

Stainless steels, hotrolled, annealed: Aluminium-base alloys:

AISI 410 135-165 C Wrought:


AISI 416 145-185 A 1100-O, annealed 23 B
AISI 430 145-185 C 1100-H, hard 44 B
AISI 446 140-185 C 3003-O, annealed 28 B
AISI 302 135-185 D 3003-H hard 55 B
AISI 303 130-150 B 2011-T3, HT and aged 95 A
AISI 345 135-185 D 2011-T8, HT and aged 100 A
AISI 316 135-185 D 2014-T, HT and aged* 130 A
AISI 309 140-185 D 2017-O, annealed 45 B
AISI 310 145-210 D 2017-T, HT and aged 100 A
Wrought and cast
nickel-base alloys: 2117-T, HT 70 A
Wrought: 0218-T, HT and aged* 100 A
Nickel (comm. pure),
Design of Cutting Tools 231

AISI no. or commo- BHN Machinabi- Aluminium Association BHN Machinabi-


nly used trade name lity rating* designation lity rating*

Stainless steels, hotrolled, annealed: Aluminium-base alloys:

annealed 100 C 2024-O, annealed 42 B


Low-carbon nickel º C 2024-T, HT 105 A
D nickel, annealed 140 C 2024-RT, HT 116 A
Z nickel, hotrolled 180 D 2025-T, HT and aged* 100 B
Monel, annealed 125 B 4032-T, HT and aged* 115 B
R Monel, hotrolled 145 A 6151-T, HT and aged* 90 B
K Monel, hotrolled 160 C 5052-O, annealed 45 B
KR Monel, hotrolled 185 B 5052-H, hard 85 B
Inconel, annealed 150 C 53S-O, annealed 26 B
Hastelloy A, hotrolled, 53S-T, HT 80 B
annealed 200-215 B 53S-W, quenched 65 B
Hastelloy B, hotrolled, 6061-O, annealed 30 B
annealed 210-235 C 6061-T, HT 95 B
Hastelloy C, hotrolled, 6061-W, quenched 65 B
annealed 160-210 C Sand-casting alloys
Illuim R, hotrolled,
annealed 190 B (except alloy K,
Sand-cast: rating C) in all
Nickel, as-cast 100 C compositions and
Monel, as-cast 135 B conditions. ... B
H Monel, as-cast 210 C Permanent-mould
S Monel, annealed 275 C casting alloys, and
Inconel, as-cast 175 C as-cast die-casting
Hastelloy A, as-cast 155-200 B alloys, in all comp-
Hastelloy B, as-cast 190-230 C ositions and cond- ... F
Hastelloy C, as-cast 175-215 C itions.
Hastelloy D º D
Illium G, as-cast 180 C
Illium R, as-cast 180 C
* A, excellent, B, good; C, fair, D, poor. * Forging grade in the indicated conditions.
SOURCE: ASTME, “Tool Engineers Handbook,” 2d ed., F. W. Wilson (ed.), McGraw-Hill, New York, 1959, by
permission.

A similar condition exists when machining low-carbon steels. Low-carbon steels


with less than about 0.15 per cent carbon machine poorly because they are soft and
gummy and adhere to the cutting tool. An increase in carbon content results in the
increase of machinablility up to a point. Thereafter, additional carbon increases the
hardness but is detrimental to machinability.
An explanation for this behaviour in carbon steels may be obtained by considering
the metallurgical aspects of iron and carbon, discussed in some detail in Ch. 3 with
regard to heat treatment of metals. The following information is similar to that in Ch.
3, the major difference being that the discussion concerns the effect of carbon on the
machinability of carbon steels.
232 Tool Design

Pure iron is soft and gummy. Under a microscope it has a white appearance, broken
only by random lines that are grain boundaries. When carbon is added, it combines to
form a compound of iron and carbon known as iron carbide. Iron carbide is hard and
brittle and very difficult to machine. Fortunately, it does not form in large lump but is
evenly dispersed throughout the parent material in the form of narrow layers of iron
carbide and nearly pure iron (ferrite). These laminated areas of cementite and ferrite
are referred to as pearlite (see Fig. 4.30). As the carbon content is increased, the
pearlite areas increase and the amount of ferrite decreases until 0.85 per cent carbon
is reached. At this point, the structure is all pearlite. When the carbon content is higher
than 0.85 per cent, and envelope of iron carbide surrounds the pearlite.

Fig. 4.30 The effect of additions of carbon to the microstructure of plain carbon steel

Thus, an increase in carbon in very low carbon steels adds iron carbides, which in
turn increases the total hardness of the material and reduces the ductility and gummy
characteristics. The result is better machinability. Additional amounts of carbon
continue to increase hardness and machinability up to about 0.20 per cent carbon.
Thereafter, additional amounts of carbon are detrimental to machinability because
the pearlitic grains (containing layers of iron carbide) become so numerous that they
interfere with the penetration of the cutting tool. A 100 per cent pearlitic structure has
a very low machinability rating for this reason.
The machinability of a cold-drawn steel is usually higher than an equivalent hot-rolled
steel. Here again the reason is that cold-drawn steel has increased hardness because
of work hardening during the cold-drawing operation. The ductility of the material is
reduced.
Heat treating can change the machinability of a material. A 100 per cent pearlitic grain
structure contains an excess of hard plates of iron carbide. The high hardness of iron
carbide requires the cutting tool to cut through the plates, i.e., fracture them, or push
between them in order to remove metal. A heat-treating process which will improve
the machinability in this condition is known as spheroidising or spheroidise annealing.
This process reduces the length of the carbide plates and produces a spheroidal or
globular from. This structure is easier to machine, as the tool can get between the
carbide more easily. Chapter 3 explains spheroidising in more detail.
The machinability of steels is often improved by the addition of free-machining addi-
tives. Steels receiving these additives are referred to as free-cutting steels. Sulphur
and lead are the most common additives. Sulphur combines with manganese in steel
Design of Cutting Tools 233

and forms sulphide inclusions in the grain Crater


structure. The inclusions may become elon-
gated by rolling at the mills. Lead inclusions
cause discontinuities in the steel structure
Wear land
and cause the chip to break as it is formed.
They also provide a “built-in lubrication sys-
tem” that reduces friction between the chip
and tool face and helps prevent chips from
sticking to the tool.
Fig. 4.31 Typical tool wear
Tool wear Cutting tools usually reach the
end of their useful life either by breaking or by wearing. Breaking is usually caused by
overloading or neglect and cannot be considered as normal wear. In general, normal
wear refers to abrasion on the flank below the cutting edge and abrasion of the tool
face just back of the cutting edge (see Fig. 4.31).
Figure 4.32 shows a two-dimensional view of a cutting tool while producing a continu-
ous chip. The areas of wear are shown and designated as zone A and zone B. Wear
zone A is at the tool-chip interface (where the chip rubs against the face of the cutting
tool). The friction between the chip and the tool causes heat, the amount depending
upon the machinability of the workpiece, the roughness of the tool, the degree of lubri-
cation, and the total contact area. The hot flowing chip may erode a groove in the tool
face back of the cutting edge when machining certain materials. This is characteristic
of machining steel and other ductile ferrous materials and is generally referred to as
cratering. Cratering is caused by a welding and galling action between the work mate-
rial and the cutting tool that tends to wash out small particles of the tool material. The
crater becomes progressively deeper until the point of the tool breaks off, resulting in
very rapid increase in temperature and total tool breakdown (see Fig. 4.33).

Wear zone A Tool Progressive Chip


(wear by cratering) Chip cratering on Tool
tool face
Wear zone B Tool
(flank wear) failure

Workpiece Built-up
Workpiece
edge

Fig. 4.32 Areas of tool wear Fig. 4.33 Tool wear by cratering

At low cutting speeds in the high-speed-steel cutting-speed range, a built-up edge


may form. The crater will start next to the BUE. Some authorities claim the BUE has
a beneficial effect; in that it tends to protect the point of the tool against the high tem-
perature in wear zone A. It has the effect of shifting the maximum temperature from
the cutting edge, which causes wear to take the form of a crater. Other authorities
claim that although the BUE is generally considered to be stagnant, it does slough off
from time to time and may remove parts of the cutting edge with it to cause a minute
chip. Repeated chipping will cause excessive wear, and thus the BUE may be detri-
mental at slow cutting speeds.
234 Tool Design

Abrasion in wear zone B results in what is known as flank wear. The amount of flank
wear is determined by measuring the width of the wear land (see Fig. 4.31). Generally
speaking, 0.8 mm flank wear is considered to be the maximum allowable wear land.
In other words, the cutting tool is considered dull when it shows a 0.8 mm wear land
below the cutting edge. A wear land in excess of 0.8 mm rapidly increase tool pres-
sure and tool wear and will result in complete tool failure by breakage. When a tool is
to be resharpened, a wear land greater than 0.8 mm rapidly decreases the number
of possible regrinds because of the extra volume of cutting-tool material that must be
removed.
When surface finish or workpiece size are of importance, a wear land of less than 0.8
mm may be the end point. Fragile workpiece may require a smaller wear land as the
forces caused by greater wear may spring the workpiece out of line. Work material
that is susceptible to work hardening cannot be machined with the maximum wear
land because of excessive rubbing between the work material and the cutting edge.
Small cutting tools that have a tendency to spring (small end mill, boring bars, etc.)
cannot be retained in services with maximum wear lands because excessive cutting
forces tend to push them away from the workpiece.
Thus, the allowable wear land depends upon many factors such as surface finish,
work materials, tool material, rigidity of setup, etc. The range of wear land width usu-
ally varies from 0.013 to 0.8 mm depending upon the above factors. Some companies
have used indexable throwaway carbide inserts until 1.5 mm wear land was obtained,
but this is possible only with heavy machining.
Flank wear is characteristic when cutting nonferrous metals or ferrous metals with
discontinuous or crumbly chips. It is also characteristic of light finishing cuts in steel.
It is commonly used as the criterion of tool-life studies because it is easy to measure
and observe. Common shop practise is to measure flank wear with a scale gradu-
ated in hundredths, but formal tool-life studies require measurement by toolmaker’s
microscope with a graduated stage. Flank wear can be predicted after a wear rate
has been established through tool-life studies because there is usually a straight-line
relationship between time in the cut and wear-land width after an initial rapid wear
period (see Fig. 4.34).

Fig. 4.34 Rate of tool wear (From R G Brierley and H J Siekmann, “Machining Principles
and Cost Control,” McGraw-Hill, New York, 1964, by permission)
Design of Cutting Tools 235

Cratering and flank wear describe tool wear in general terms. A third factor contribut-
ing to tool wear is mechanical wear due to breaking out of small chips from the cutting
edge. This is characteristic of brittle cutting-tool materials such as carbides of oxides.
This type of wear is usually due to mechanical or thermal shock caused by vibration
and impact of the machine and alternate heating and cooling of the cutting too. Minute
chipping may occur in the life of the tool because of a fine and fragile cutting edge
but may be disguised or covered up by flank wear. Chipping of this type may cause
premature flank wear and result in shorter overall tool life.
The mechanisms causing tool wear is varied and many are not thoroughly understood.
However, one factor is recognised as contributing to tool wear more than any other.
This is the temperature at which the material is machined. The temperature depends
upon the cutting speed. When the cutting speed is increased, so is the temperature.
It is said that over 95 per cent of the power used in machining is converted to heat.
Heat is a measure of the amount of energy consumed. The same amount of energy is
required to remove a given amount of material, regardless of cutting speed.
At extremely slow cutting speeds, the heat is carried off in the chip and through the
workpiece, cutting tool, and atmosphere. The level of energy, or temperature, remains
low in the area of the cutting-tool edge. Temperature at this speed is not a problem
with modern cutting-tool materials. However, when the cutting speed is increased,
the level of energy or temperature is increased sharply, as there is not enough time
for heat dissipation. A point is reached where temperature begins to affect tool life,
depending upon the cutting-tool material. Unfortunately, this point is well below the
speed capabilities of modern machine tools.
Today’s production requirements prohibit the machining of metal at this low speed.
Therefore, cutting speed is increased and temperature, or the level of energy, becomes
the major factor affecting tool life. Some cutting-tool materials withstand higher tem-
peratures than others, but in all cases tool life is shortened as the cutting speed is
increased. In general, this is due to the influence of temperature on the hardness of
tool materials. Figure 4.35 shows the relationship of temperature to the hardness of
cutting-tool materials.

Fig. 4.35 Relationship between temperature and the hardness of cutting-tool material
(Metallurgical Product Department, General Electric Company)
236 Tool Design

Heat is generated in two main areas. The first is the area of shear, where heating is
caused by deformation of the metal. The bulk of the heat produced is in this area. The
second area of heat is produced by friction where the chips rub across the face of
the tool. The temperature here is determined by the area of contact and the frictional
behaviour of the work and tool material. Higher cutting speeds cause a temperature
rise in both areas, but temperature in the tool-chip interface area has the greatest
effect on tool life.
Paradoxical as it may seem, the amount of heat that remains in the cutting tool or fin-
ished workpiece is reduced as the cutting speed is increased. Figure 4.36 shows the
distribution of heat in work, tool, and chips at various speeds. At high speeds, the chip
travels so fast that most of the heat is carried off with it. When machining at extremely
high speeds with oxide cutting tools, it is possible to stop the machine and touch the
workpiece and cutting-tool insert without getting burned fingers, although the chips
are coming off at a dull-red heat.

Fig. 4.36 Distribution of heat in work, tool, and chips at various cutting speeds. (From R G
Brierley and H J Siekmann, “Machining Principles and Cost Control,” McGraw-
Hill, New York, 1964, by permission)

The above discussion should not imply that a high temperature is not reached at the
tool-chip interface. At very high speeds the tool-chip interface temperature is quite
high, as shown in Fig. 4.37. It reaches a maximum about two-thirds of the way back
from the cutting edge. This is the region where the crater is formed on the face of the
cutting tool. As the chip slides over this area at increasing speed, the temperature
rises to a point where atomic diffusion between the two materials takes place. One
theory is that when diffusion take place, an alloy of the two materials is formed, which
in turn will lower the melting point and weaken a thin film of tool material. Particles
of the tool material are swept away with the chip, and thus the crater is formed. This
theory is supported by the observation that once the temperature exceeds a certain
value, crater wear increases quite rapidly.
Once a flank wear land is started by abrasion, frictional heat is produced in this area.
It is believed by some authorities that the temperature may reach the diffusion point
as the wear land becomes wider and friction increases. This would suggest that the
Design of Cutting Tools 237

Fig. 4.37 Relationship of tool-chip interface temperature to cutting speed


wear mechanism on the tool flank is influenced by temperature in a manner similar
to crater wear. Thus, changing to a cutting-tool material that requires a higher tem-
perature for diffusion should reduce both cratering and flank wear when machining
at high speeds.
Cutting speed has the greatest influence on tool temperatures. Feed also has an
effect on temperatures, but not as great as speed. Figure 4.38 shows the relation-
ship of cutting temperature to feed. The depth of cut has little effect on temperature
because a greater area of the tool is used which will dissipate additional heat. When
a large amount of metal is to be removed and temperature is a prime consideration,
it would be better to increase the feed rather than the speeds, provided that the tool
and workpiece are rigid enough to withstand additional feed.

Fig. 4.38 The relationship of cutting temperature to feed


238 Tool Design

Speeds and feeds In the past, the selection of cutting speeds and feeds has gener-
ally been left to the machinist or shop foreman. The choice was based on cut-and-try
methods, and the results in most cases have been satisfactory. Cut-and-try methods
will always be followed to a certain extent, but more and more speeds and feeds are
being determined by methods men from the engineering department. There are sev-
eral reasons for following this practise:
1. Older machine operators tend to be conservative in their selection of feeds
and speeds in order to make their tools last longer. There was a time when
a machinist’s ability was judged by how long he could make a tool last
between regrinds. A tool change could be a major job with considerable
machine downtime and lost production. However, with the introduction of
throwaway insert-type tools, it is no longer necessary for the cutting tool to
last for extended periods of time. A new cutting edge can be indexed in a
matter of seconds without shifting the position of the toolholder. The opera-
tor is able to change to a new cutting edge with practically no lost produc-
tion time. Yet older operators still tend to operate machines at speeds and
conditions to get maximum tool life in order to minimise machine downtime.
This tendency is reduced when speed and feeds are determined elsewhere.
2. Highly skilled operators who have the ability to determine optimum feeds and
speeds are not readily available. It is often necessary to hire operators with
very little experience or training. Such a condition forces management to des-
ignate an individual or department to determine the proper feeds and speeds.
3. Management retains a greater amount of control in production rates. Once time
studies have been made at known optimum feeds and speeds, management
can be relatively sure of subsequent production rates. Estimating is also more
accurate if known feeds and speeds are available. This would not be possible
if each operator determined feeds and speeds to suit his own taste.
Since engineering is assuming a greater role in specifying feeds and speeds, the
tool designer should have a thorough understanding of this subject. This cannot be
gained by simply referring to tables or plugging values into formulas. Tables and for-
mulas should always be considered as general recommendations. The experienced
process engineer will tell you that the selection of feeds and speeds is generally a
compromise between several variables. The determination of the proper compromise
will be possible only after a considerable amount of practical experience, but a basic
understanding will be of great help.
The cutting speed at which a material is machined depends upon many variable fac-
tors. The machinability of the work material, the tool geometry, the type of lubricant or
coolant, the feed and depth of cut, and the power and design of the machine tool are
generally the most influential factors.
In machining steel, the allowable cutting for a given tool life between grinding or index-
ing tool inserts is, as a general rule, inversely proportional to the hardness of the given
steel. The machinability of other steels of the same hardness may vary, however, as
steel with different allying elements may be the same hardness but each one may
have more or less resistance to cutting. One may be machined faster than the other,
even though they have the same hardness.
Cutting speed is the distance travelled by the work surface in unit time with reference
to the cutting edge of the tool. The cutting speed is usually expressed in m/min.
Design of Cutting Tools 239

The cutting speed at which a meterial is machined depends upon many variable
factors. The machinability of the work material, the tool geometry, the type of lubricant
or coolant, the feed and dept of cut, and the power and design of the machine tool are
generally the most influential factors.
In machining steel, the allowable cutting for a given tool life between grinding or
indexing tool inserts is, as a general rule, inversely proportional to the hardness of
the given steel. The machinability of other steels of the same hardness may vary,
however, as steel with different alloying elements may be the same hardness buy
each one may have more or less resistance to cutting. One may be machined faster
than the other, even though they have the same hardness.
The correct surface speed for machining various materials is determined form differ-
ent sources. Many machine operators rely entirely on past experiences when select-
ing speeds and have no idea at what cutting speed the material is being machined.
Using cut-and-try methods for establishing speed rates may have been satisfactory in
the days when few new materials were encountered, but they have no place in mod-
ern industry, where new materials are encountered everyday. Too much time would
be spent in cutting and trying.
A rule-of-thumb method for determining cutting speed used by many works well for
determining general speeds and as a starting guide when speed charts are not avail-
able. The rule of thumb applies to high-speed cutting tools and is as follows:
1. Remember that cold-rolled or mild-steel will be machined in the neighbourhood
of 30 m/min
2. To machine the tougher materials, the machining speed of mild steel will be
cut in half. Thus, the machining speed for tough materials will be 50 m/min.
(Tough steels include tool steels, drill rod, chrome alloys, nickel alloys, vana-
dium alloys, etc.)
3. To machine the softer materials, the machining speed of mild steel will be dou-
bled. Thus, the machining speed of soft materials will be 61 m/min. (Aluminium
is an exception to the rule: it should be machined from three to four times as
fact.)
Again, these speeds are only approximate and are used when the material is being
machined with high speed-steel cutting tools. They are for broad, general use and are
subject to change because of the variables occurring in tools, materials, and set-up.
When machining with cast-alloy tools, these high-speed steel speeds may be dou-
bled. Mild steel should then be machined at approximately 61 m/min when using
cast-alloy cutting tools. The HSS speeds may be tripled when using carbide cutting
tools, and 91 m/min would then be a good starting point for mild steel when machining
with carbides. A good general rule for oxide cutting tools is to make initial test cuts of
between two and three times the speeds normally used for carbides.
This rule of thumb is useful in that it is easy to remember and enables the beginner to
have some idea of the correct speed for various materials. It specifies cutting speeds
in wide general terms, but it is still much better than a guess.
Reference to handbook speed tables and cutting-tool manufacturer’s recommenda-
tions is a more exact method of determining cutting speeds, especially when machin-
ing a new or unfamiliar material. Here again, table and handbook recommendations
must be considered as general because of the many variable factors in machining.
Tables 4.7 to 4.9 are typical examples of speed tables.
Table 4.7 Cutting speeds and feeds for turning steels

Steel Condition BHN High-speed steel Cast alloy Carbide Oxide


Speed, Feed, Speed, Feed, Speed, Feed, Speed, Feed,
m/mm mm/rev m/mm mm/rev m/mm mm/rev m/mm mm/rev
240 Tool Design

Low-carbon, Cold-drawn 170-190 58 0.31 76.2 0.31 190 0.38 183-457 0.13-0.51
free-machining.
Medium-carbon, Cold-drawn 200-230 43 0.31 61 0.31 128 0.31 137-381 0.13-0.51
free-machining.
Quenched 250-300 29 0.31 37 0.31 122 0.31 127-305 0.13-0.38
tempered and
Plain low-carbonv Annealed 110-165 43 0.31 55 0.31 160 0.38 168-457 0.13-0.51
Plain medium- Annealed 120-185 30 0.31 46 0.31 145 0.38 137-305 0.13-0.38
carbon (0.40 to
0.50 C)
Plain high-carbon Annealed 170-200 27 0.31 43 0.31 130 0.31 130-274 0.13-0.38
(0.55 to 0.95 C) Quenched 210-250 24 0.25 38 0.25 122 0.31 122-244 0.13-0.38
Plain medium-carbon and tempered
Plain medium Quenched and 260-310 21 0.25 34 0.25 107 0.25 99-229 0.13-0.38
carbon tempered
Plain high-carbon Quenched and 320-375 15 0.25 21 0.25 69 0.25 76-213 0.13-0.38
tempered
Resulphurised alloy Annealed 160-210 38 0.31 53 0.31 130 0.38 137-274 0.13-0.38
Leaded alloy Annealed 140-190 46 0.31 69 0.31 145 0.38 183-610 0.13-0.51
Normalised 250-300 24 0.25 34 0.25 122 0.31 137-305 0.13-0.51

Contd...
Contd...

Steel Condition BHN High-speed steel Cast alloy Carbide Oxide


Speed, Feed, Speed, Feed, Speed, Feed, Speed, Feed,
m/mm mm/rev m/mm mm/rev m/mm mm/rev m/mm mm/rev

Alloy steels Annealed 150-240 24-34 0.25 32-38 0.25 91-130 0.38 107-305 0.13-0.51
Normalised or 240-310 20 0.20 26 0.25 99 0.31 91-305 0.13-0.38
quenched and
tempered
Quenched and 315-370 14 0.20 20 0.25 84 0.25 91-274 0.13-0.38
tempered
Quenched and 380-440 11 0.20 17 0.25 76 0.25 76-244 0.13-0.31
tempered
Quenched and 450-500 8 0.20 14 0.25 55 0.25 76-213 0.13-0.25
tempered
Quenched and 510-560 5 0.20 8 0.25 37 0.25 61-183 0.08-0.2
tempered

SOURCE: ASTME, “Tool Engineers Handbook,” 2d ed., F W Wilson (ed.), McGraw-Hill, New York, 1959, by permission.
Design of Cutting Tools 241
242 Tool Design

Table 4.8 Cutting speeds and feeds for turning nonferrous materials
Work material Condition BHN* High-speed steel Cast alloy Carbide
Speed, Feed, Speed, Feed, Speed, Feed,
m/min mm/rev m/min mm/rev m/min mm/rev
Aluminium alloys

Nonheat-treatable Cast 50-70 274 0.31 366 0.31 Max 0.38


cast alloys
Heat-treatable Solution- 70-105 213 0.31 305 0.31 Max 0.38
cast alloys treated and
aged
Nonheat-treatable
wrought alloys Cold-drawn 40-70 213 0.31 366 0.31 Max 0.38
Heat-treatable Solution- 65-105 213 0.31 305 0.31 Max 0.38
wrought alloys treated and
aged
Magnesium alloys
Cast alloys:
A10, A12, AZ63,
AZ63X, AZ101,
AZ92, AZ92X,
AS100, AZ90,
AZ90X Cast 35-70 274 0.31 457 0.31 Max 0.31
Wrought alloys:
AT35, AZ31,
AZ61, AZ80 Cold-drawn 40-80
274 0.31 457 0.31 Max 0.31
Copper alloys
Group I* Wrought or 120-160 122 0.25 198 0.25 305 0.25
cast
Group II* Wrought or 165-180 84 0.25 152 0.25 229 0.25
cast
Group III* Wrought or 172-205 38 0.25 99 0.25 183 0.25
cast
Titanium alloys
Commercially Wrought or 150-200 46 0.25 49 0.25 114 0.25
pure cast
Alloys:
MST 5A1-2.5Sn, Wrought or 250-320 11 0.25 15 0.25 46 0.25
Rs 110C, cast
A110AT, Ti
6A1-4Zr-1V,
Ti 8A11Mo-IV
Alloys:
Ti6A1-4V, Ti Wrought or Over 320 5 0.25 8 0.25 38 0.25
2A1-16V, Ti cast
4a1-4Mo-4V
* 500kg load for aluminium and magnesium.
* Group I: Free-cutting yellow brass, clock brass, low leaded brass, forging brass, free-cutting tube brass, medium leaded brass, leaded
naval brass, 88-4-4-4 bronze, hardware bronze, leaded bronze, leaded silicon bronze, leaded commercial bronze, tellurium copper,
selenium copper, and leaded copper.
* Group II: Leaded muntz metal, leaded copper silicon alloys, muntz metal, copper silicon alloys, admiralty metal, naval brass, car-
tridge brass, yellow brass, red brass, leaded phosphor bronze, manganese bronze, tobin bronze silicon bronze, and leaded nickel silver
(12%).
* Group III: Nickel silver (18%), cupronickel, commercial bronze, phosphor bronze, nickel-aluminium bronze, beryllium copper, chro-
mium copper, and copper.
SOURCE: ASTME, “Tool Engineers Handbook,” 2d ed., FW Wilson (ed.), McGraw-Hill, New York, 1959, by permission.
Design of Cutting Tools 243

Lower feeds are used for light finishing cuts, frail set-ups, hard-to-machine work mate-
rials, frail and small cutters, and for milling deep slots.
The feed of a lathe and similar machine tools is the distance the tool is fed for each
revolution of the work. It is expressed in mm/rev. Lathe feeds may vary from 0.07 to
1.8 mm/rev depending upon the size of the lathe, the machinability of the material, the
finish required, the setup of the workpiece, etc. Tables 4.7 and 4.8 give various feeds
for turning operations.
The feed of a drill is the rate at which the drill is advanced into the workpiece per revo-
lution of the drill. Like that of the lathe, it is usually expressed in mm/rev. The feed is
governed by the size of the drill, the type of material being drilled, and the rpm of the
drill. Smaller drills require less feed than larger drills. Too slow a feed rate may retard
production rates, but too heavy a feed rate may cause a drill to split up the web. Drill
feed ranges usually vary from 0.025 to 0.6 mm/rev for drills ranging from 1.6 to 25 mm
in diameter. Table 4.10 gives feed rates for various drilling operations.
The feeds for reaming are usually higher than those used for drilling because ream-
ers have more teeth and a reaming operation is for sizing and finishing. Since reamer
tooth does not have to remove a lot of material, the chip load per tooth may be higher.
There should be sufficient material left for reaming, as insufficient material may result
in a burnishing rather than a cutting action. The amount of material to be removed by
a reamer depends upon the type of material, feed, finish required, and the size of the
reamer. Generally speaking, when using a machine reamer,0.25 mm of material is left
for a 6 mm hole, 0.4 mm for a 13 mm hole, 0.5 mm for a 19 mm hole, 0.6 mm for a 25
mm hole, and 0.75 mm for a 32 mm hole. Material allowances are much smaller for a
hand reamer because of the difficulty in forcing the reamer into the material. Common
allowances for hand reamers are from 0.025 to 0.08 mm.
A good starting point for reamer feeds is from 0.025 to 0.10 mm per flute per revolu-
tion. A low feed may result in glazing, tooth rubbing, or chatter. High feed rates tend
to reduce the accuracy of the hole and quality of the surface finish. Feeds and speeds
for high-speed steel reamers are given in Table 4.11.
The feed of shapers and planers is the distance the cutting tool advances at the
end of each cutting stroke. Thus, if the feed is set at 0.6 mm the tool will move over
0.6 mm at the end of the cutting stroke, after the tool clears the workpiece. Feeds for
shapers vary from 0.25 to 4.3 mm per double stroke. Feeds for planers are consider-
ably heavier, and with a special ground tool, the planer feed may be as high as 19
mm; however, for general work the feeds vary from 0.8 to 3.2 mm. Table 4.12 gives
feeds for shapers and planers when machining various materials.
Since the feed and speed of a milling machine are independent of each other, estab-
lishing the feed is more difficult. The feed of a milling machine is expressed in mm per
minute of table movement rather than mm per spindle rotation. If the feed of a milling
machine is set at 76 mm, the mill table will travel 76 mm in 1 min. Generally, the feeds
on a milling machine vary from 13 to 2542 mm per min on large machines. Feeds on
smaller knee and column mills vary from 13 to 762 mm per min.
The feed in a milling machine is defined as the rate with which the workpiece advances
under the cutter. The feed is expressed in a milling by the following three methods:
244 Tool Design

Table 4.9 Cutting speeds for milling, m/min


Work material High-speed steel tools Carbide-tipped tools Cutting fluid
Rough mill Finish mill Rough mill Finish mill
Cast iron 15-18 24-34 55-61 107-122 Dry
Semisteel 12-15 20-27 43-49 76-91 Dry
Malleable iron 29-30 34-40 76-91 122-152 Soluble, sulphurised,
or mineral oil
Cast steel 14-8 21-27 46-55 61-76 Soluble, sulphurised,
mineral or mineral-
lard oil
Copper 30-46 46-61 183 305 Soluble, sulphurised
or mineral-lard oil
Brass 61-91 61-91 183-305 183-305 Dry
Bronze 30-46 46-55 183 305 Soluble, sulphurised
oil, or mineral-lard oil
Aluminium 122 213 244 Soluble, sulphurised
oil, mineral oil and
kerosine
Magnesium 183-244 305-457 305-457 305 Dry, kerosine or
mineral-
lard oil
SAE steels:
1020 (course 18-24 18-24 61 61 Soluble, sulphurised,
feed mineral or mineral-
lard oil
1020 (fine feed) 30-37 30-27 137 137
1035 23-27 27-37 76 76
X-1315 53-61 53-61 122-152 400-500
1050 18-24 30 61 61
2315 27-34 27-34 91 91
3150 15-18 21-27 61 61
4150 12-15 21-27 61 61
4340 12-15 18-21 61 61 Sulphurised and
mineral oil
Stainless steel 18-24 30-37 73-91 73-91
Titanium 9-21 61-107

NOTE: Feeds should be as much as the work and equipment will stand, provided a satisfactory surface finish is
obtained.
SOURCE: ASTME, “Tool Engineers Handbook,” 2d ed., F W Wilson (ed.), McGraw-Hill, New York, 1959, by
permission.

Feed per tooth (fz) The feed per tooth is defined by the distance the work advances
in time between engagements by the successive teeth. It is expressed in millimeters
per tooth of the cutter.
Feed per cutter revolution (fr) The feed per cutter revolution is defined by the
distance the work advances in time when the cutter turns through one complete
revolution. It is expressed in millimeters per revolution of the cutter.
Design of Cutting Tools 245

Table 4.10 Speeds and feeds used for drilling with high-speed steel drills

Material Hardness Peripheral Feed,


Brinell Rockwell speed, m/min or mm/rev

Allegheny metal 146-149 B78-80 15 Medium


Aluminium 90-101 B54-56 198-396 0.13-0.9
Aluminium bronze 170-187 B86-90 18 Medium
Armor plate* 444 C46-47 3-4 Light
Asbestos* ... ... 17 Medium
Bakelite* ... ... 24 Medium
Brake linings, etc ... ... 17 Medium
Brass* 192-202 B92-94 61-76 Heavy
Bronze, common 166-183 B85-89 61-76 Heavy
Phosphor, one-half hard 187-202 B99-94 53-55 Medium
Soft 149-163 B80-84 61-76 Heavy
Carbon electrodes ... ... 18-21 Light
Cast iron, soft 126 B70 43-46 Heavy
Medium 196 B93 24-34 Medium
hard 293-302 C23-33 14-15 Light
chilled* 402 C42-43 5 Light
Cast steel 286-302 C30-33 12-15 Light
Celluloid ... ... 30 Medium
Copper 80-50 B40-44 21 Light
Copper alloys, machinability
70-100 ... ... 61-152
30-70 ... ... 23-76 0.08-0.51
20 ... ... 15-38
Drop forgings and tool steel 170-196 B86-93 18 Medium
Duralumin 90-104 B48-58 61 Medium
Everdur 179-207 B88-95 18 Light
Formica (laminated Bakelite) ... ... 17 Light
Glass* ... ... 5 Light
Gold, white 134-149 B79-80 8 Light
Machinery steel 170-196 B86-93 34 Heavy
Magnesium, sheet* ... ... 91-610 0.13-0.76
for 6.36 mm
to 25.4 mm
diameter drills
Shallow hole ... ... 91-610 0.1-1.27
for 6.36 mm
to 25.4 mm
diameter drills
Deep hole ... ... 91-610 0.1-0.76
for 6.36 mm
to 25.4 mm
diameter drills
Magnet steel, soft C24-32 241-302 11-12 Medium
246 Tool Design

Material Hardness Peripheral Feed,


Brinell Rockwell speed, m/min or mm/rev
Hard* C34-51 321-512 5 Light
Manganese steel 7-13%* B90-96 187-217 5 Light constant
Manganese copper 30% B74 134 5 Light
Malleable iron B66-70 112-126 26-127 Heavy
Marble ... ... 5 Light
Mild steel, 20-30 carbon B86-93 170-202 34-37 Heavy
Molybdenum steel, conn. rod B92-99 196-235 17 Medium
Monel metal B80-86 149-170 15 Medium
Nickel pure* B90-94 187-202 23 Light
Nickel steel, 31/2% B93-100 196-24 1 18 Light
Nitralloy 135 (G) C22-25 228-250 15-18 Light
Permalloy, 77% nickel B72-84 131-163 15 Light
Rubber, hard* ... ... 30 Light
Screw stock (cold-rolled) B86-93 170-196 34 Heavy
Slate* ... ... 5 Light
Spring steel C43 402 6 Light
Stainless steel B78-80 146-149 15 Medium
Cold-rolled C48-49 460-477 6 Light
Steel, 40-50 carbon B86-93 170-196 24 Medium
Titanium* ... ... 9-21 0.08-0.2
Tool steel SAE and B80 149 23 Heavy
forging steel B93 196 18 Medium
C24 214 15 Medium
C32 302* 12 Light
C38 351 6 Light
C43 402* 5 Light
Wood, hard ... ... 152 Light
Zamak (zinc alloy)* B62-70 112-126 61-76 Medium

* Key to feed per revolution:

Drill, mm Heavy feed, mm/rev Medium feed mm/rev Light

6.36 0.2 0.13 0.06


12.71 0.25 0.19 0.10
19.06 0.41 0.25 0.15
25.4 0.51 0.31 0.19
* Use cobalt drills.
* Use tungsten carbide-tipped drill.
* Brooks & Perkins.
* Rem-Cru Titanium, Inc.
SOURCE: ASTME, “Tool Engineers Handbook,” 2d ed., F W Wilson (ed.), McGraw-Hill, New York, 1959, by
permission.
Design of Cutting Tools 247

Table 4.11 Speeds and feeds for high-speed-steel reamers


Material Speed, m/min Feed, mm/rev
Aluminium 50 A
Bakelite 21 A
Brass, leaded 55 A
Red or yellow 46 A
Bronze, cast 46 B
Soft 55 A
Copper 13 C
Duralumin 37 A
Everdur 11 B
Glass 13 D
Cast iron, chilled 13 D
Hard 23 C
Medium 27 A
Pearlite 18 A
Soft 34 A
Malleable iron 20 B
Magnesium 67 A
Monel 9 C
Nickel 13 C
Plastic 23 C
Rubber, hard 18 B
Steel, BHN 130 29 A
150 25 B
170 20 B
200 17 B
230 15 C
260 12 C
300 9 C
330 6 C
360 5 D
400 3 D
Cast 7 C
Forged alloy 9 C
Forged carbon 11 C
Low-carbon 22 B
Medium-carbon 20 B or C
High-carbon 14 D
Stainless 4 C
Tool 11 D
Titanium 12 A
Wood, hard 91 A
Zinc alloy 46 A
* Up to 150% higher for carbide reamers.
248 Tool Design

Reamer mm. A B C D
3.18 0.15 0.13 0.10 0.08
12.71 0.31 0.25 0.18 0.13
25.4 0.51 0.41 0.31 0.20
50.84 0.81 0.66 0.51 0.33
63.55 1.09 0.89 0.71 0.46
76.26 1.42 1.14 0.89 0.58

SOURCE: ASTME, “Manufacturing Planning and Estimating Handbook,” F. W. Wilson (ed.), McGraw-Hill, New
York, 1963, by permission.

Table 4.12 Speeds and feeds for shapers and planers


Work material Type of tool
High-speed steel Alloy high speed Cast alloys Carbides
Speed, Max Speed, Max Speed, Max Speed, Max
feed, feed, feed, feed,
m/min mm/rev m/min mm/rev m/min mm/rev m/min mm/rev

Aluminium 61-91 Deter. * * * * Max of mach. 4.77


by finish

Brass, soft 48-76 6.36 * * * * Max of mach. 4.77

Bronze, medium 23-38 1-9 * * * * 46-91 1.27

Hard 9-18 1.27 15-21 1.27 15-30 1.02 46-61 1.27

Cast iron, soft 15-24 3.18 18-30 3.18 27-37 1.27 34-69 1.27

Hard 9-15 1.53 12-18 1-53 15-24 1.27 30-61 1.27

Malleable iron 15-27 2.29 21-34 2.29 24-37 1.27 46-76 1.27

Cast steel 8-18 1.27 12-21 1.27 18-24 1.02 30-55 1.02

(30 carbon)

Steel, soft 21-30 1.27 24-37 1.27 * * 55-91 1.27

Medium 18-21 1.53 21-27 1.53 * * 55-76 1.27

Hard 6-11 0.89 9-15 0.89 * * 30-55 0.89


NOTE: Data base on an average depth of cut 12.7 mm Speed increases up to 50 per cent are frequently possible
on light finishing cuts.
* This tool not recommended for this application.
SOURCE: ASTME, “Manufacturing Planning and Estimating Handbook,” F. W. Wilson (ed.), McGraw-Hill, New
York, 1963, by permission.

Feed per minute (fm) The feed per minute or feed rate is defined by the distance the
work advances in one minute. It is expressed in millimeters per minute.
Design of Cutting Tools 249

fm = f r × N
f m = fz × z × N

where Z = number of teeth or the cutter


N = rpm of the cutter

Manufactures of cutting tools provide feed-per-tooth tables similar to Tables 4.13.


These recommendations are intended to be general and may be used as a starting
point.

Table 4.13 Suggested feed in mm per tooth for high-speed-steel milling cutters*
Material Face Helical Slotting and End Form-relieved Circular
mills mills side mills mills cutters saws

Plastic 0.33 0.25 0.2 0.18 0.10 0.08


Magnesium and alloys 0.56 0.46 0.33 0.28 0.18 0.13
Aluminium and alloys 0.56 0.46 0.33 0.28 0.18 0.13
Brasses and bronzes, free-cutting 0.56 0.46 0.33 0.28 0.18 0.13
Medium 0.36 0.28 0.2 0.18 0.18 0.08
Hard 0.23 0.18 0.15 0.13 0.08 0.05
Copper 0.31 0.25 0.18 0.15 0.10 0.08
Cast iron, soft (150-180 BHN) 0.41 0.33 0.23 0.02 0.13 0.10
Medium (180-220 BHN) 0.33 0.25 0.18 0.18 0.10 0.08
Hard (220-300 BHN) 0.28 0.2 0.15 0.15 0.08 0.08
Malleable iron 0.31 0.25 0.18 0.15 0.10 0.08
Cast steel 0.31 0.25 0.18 0.15 0.10 0.08
Low-carbon steel, free-machining 0.31 0.25 0.18 0.15 0.10 0.08
Low-carbon steel 0.25 0.2 0.15 0.13 0.08 0.08
Medium-carbon steel 0.25 0.2 0.15 0.13 0.08 0.08
Alloy steel, annealed (180-220 BHN) 0.2 0.18 0.13 0.10 0.08 0.08
Tough (220-300 BHN) 0.15 0.13 0.10 0.08 0.05 0.05
Hard (300-400 BHN) 0.10 0.08 0.08 0.05 0.05 0.03
Stainless steels, free-machining 0.25 0.2 0.15 0.13 0.08 0.05
Stainless steels 0.15 0.13 0.10 0.08 0.05 0.05
Monel metals 0.2 0.18 0.13 0.10 0.08 0.05
Titanium 0.2 0.18 0.13 0.10 0.08 0.05
* When carbon-steel cutters are used, values in this table should be divided by 2. Data from The Cincinnati Milling
Machine Co.
SOURCE: ASTME, “Manufacturing Planning and Estimating Handbook,” F W Wilson (ed.), McGraw-Hill, New
York, 1963, by permission.

Efficiency of machining The majority of discussions concerning tool-life studies,


cutting-tool materials, throwaway inserts, the newer machine tools, etc., appear to
emphasise cutting speed. Beginner may get the impression that cutting speed is the
all-important factor, with depth of cut and feed as secondary factors. Even experi-
enced machinists tend automatically to set cutting speed first, followed by feed rate
and then depth of cut.
250 Tool Design

It has already been shown that there is a relationship between cutting speed, depth of
cut, and feed for a specific tool life when cutting a given material. Cutting speed has
the greatest influence on tool temperatures, followed by feed and depth of cut.
Depth of cut has the greatest influence upon cutting forces, followed by feed, with cut-
ting speed having the least influence. Feed has the greatest effect on surface finish
when it is set according to nose radius. With these relationships, which is the correct
compromise for the most efficient removal of material with the best possible tool life?
The goal should be to remove the greatest amount of material in the shortest length
of time, consistent with finish requirements, work and tool rigidity, available power of
the machine, and relative cost of labour and cutting tools.
The available power of the machine limits the depth of cut. Since feed has less influ-
ence on the depth of cut (see Figs. 4.23 and 4.24), it is possible to remove metal
more efficiently at a heavy feed since less power will be required. Since cutting force
is relatively unaffected by changes in speed above 30 m/min (see Fig. 4.22), a slower
speed will give more efficient use of available power. Thus, heavier feeds to obtain
good material removal rates and slower speeds to utilise available power better com-
bine to reduce the total energy requirements and improve the efficiency of the cutting
process.
Materials that strain (work) harden may be machined more efficiently with increased
feeds. As a tool dulls, the wear land rubs the finished surface and the surface layer
becomes workhardened. The following cutting edge must cut in the hardened layer
when light feeds are used. Heavier feeds transfer much of the cutting to material
below the hardened layer.
Materials having extreme abrasive properties but requiring low cutting forces can be
machined more efficiently with higher feeds. Certain plastics are examples. In this
case, the cutter makes fewer turns to complete a given surface area. If the feed is
slowed down, a greater number of turns is required to cover the area. The more turns
the cutter makes, the more wear on the cutting edges.
There are, of course, limitations on the amount of increase in cutting feed. The rate of
feed determines surface finish. An increase in feed also increases the cutting forces
on the tool and workpiece. At some point these forces will exceed the strength of tool
or workpiece, and breakage of deflection will occur.
This discussion still has not given a sound solution to the most efficient combina-
tion of feeds, speeds, and depth of cut. It suggests that metal can be removed more
efficiently by slowing down speed rates. Cutting speed is not the all-important factor.
A general rule used by many production people to achieve the greatest machining
efficiency is to use the heaviest feed that will allow the required surface finish, use the
maximum depth of cut consistent with available power and rigidity of workpiece and
machine, and then establish the cutting speed to give the desired tool life. The ques-
tion now becomes: What is the most economical tool life? Speeds may be increased
to the point where the cutting edge will last only a few minutes or decreased to the
point where the cutting edge will last hours. Too fast a speed will increase tool costs
and downtime for tool changing. Too slow a speed simply cannot produce enough
pieces to make profit. Somewhere between the “too slow” and “too fast” is a cutting
speed that will give the best tool life for overall efficiency.
Usually, the best cutting speed is the one that will reduce the total cost of machining to
a minimum cost per piece. However, cost may be secondary, and the objective may
be to set the maximum production rate.
Design of Cutting Tools 251

Many variables determine the minimum cost and maximum production rate. The cost
of labour on the machines, overhead costs, set-up time, tool costs, tool-changing
time, time to machine the workpiece, tool-grinding-room labour costs, etc. – all these
enter into the picture. The majority of these variables may be changed to known
quantities for a particular job. When these quantities are determined, it is possible to
plot costs and production rates vs. cutting speeds. The resulting graph will show the
minimum cost and the maximum production rate.
For example, if production costs are plotted against cutting speed, as shown in Fig.
4.39, the cost decreases as cutting speed increases because more workpieces
are being produced. However, a point is reached where production cost begins to
increase with additional increases in speed because of shorter tool life. This point
is the minimum cost speed. If production rate is plotted against cutting speed on the
same graph, it can be seen that as cutting speed increases, the production rate also
increases to a point of maximum production and then decreases rapidly. Note that the
speed for maximum production is faster than the speed that produces minimum cost
per piece. In maximum production is desired, one cutting speed is indicated, but if the
lowest unit machining cost is desired, a lower speed must be selected. In any case,
the theoretical “best” cutting speed will lie between the points of minimum cost and
maximum production. It is still necessary to use an element of individual judgment in
selecting the speed for the machine, but at least the tool or production engineer has
an idea where he stands.

Theoretical best
cutting speed

Maximum
production

Production Production
Rate
Cost
Minimum
cost

Cutting speed

Fig. 4.39 Graph of maximum production rate and minimum production costs

This procedure may appear to be complicated and involved. In reality, it is not. Various
cutting-tool manufacturers have simplified the procedure and developed practical
methods of determining the most efficient cutting speed for any production machining
operation.1

1 Probably the best-known method is available from the Metallurgical Products Department, General Electric
Company, who call it “hi-efficiency machining” (HI-E). It is based on minimum cost per piece and maximum
production rates, using sound engineering principles that have long been well understood and accepted by
the tool-engineering profession. The HI-E kit consists of a slide rule for calculations and a machinability rating
table. It is beyond the scope of this text to go into HI-E machining in detail, but information may be obtained
from the Metallurgical Products Department of General Electric.
252 Tool Design

Surface finish The removal of metal by an edged cutting tool generally requires one
or more roughing cuts to remove the excess metal and then a finishing cut to bring the
workpiece to a given size and establish the correct surface roughness. The average
machine operator, left to his own initiative, will usually increase the cutting speed and
reduce the feed on a finishing cut. He knows that he can take a light cut at a greater
speed than a heavy cut and that a slower feed will cause the feed marks to be closer
together. This is usually the extent of his understanding of surface finish and its rela-
tion to the other machining variables. The purpose of this discussion is to examine
these machining variables and their effect on surface finish.
Irregularities produced by the cutting tool are referred to as surface roughness. Three
main factors determine surface roughness: the rate of feed, the built-up edge, and
vibrational deflections between the tool and the work. The rate of feed controls the
magnitude of the feed marks produced by the cutting tool. The built-up edge sheds
fragments of the chip on the surface and is responsible for a “fuzzy” appearance.
Vibrational deflection, generally referred to as chatter, it is condition of sustained
vibration in which the cutting tool and workpiece shake at a definite frequency. This
vibration is primarily caused by the variation in the cutting force caused by pressure
being built up on the chip and then dropping as the chip separates from the work-
piece. The result is a rhythmic pattern on the surface commonly referred to as chatter
marks. For further discussion on chatter, see page 166.
The rate of feed is one of the more important considerations in obtaining the desired
surface roughness. The average machine operator reasons that the feed marks are
closer together when the feed rate is reduced. This is true, but the real consideration
is the relationship between the feed rate and the nose radius of the tool. A feed rate
that is equal the nose radius would give a poor finish, as shown in Fig. 4.40. However,
if the nose radius is increased, as shown in Fig. 4.40b, the feed marks tend to be
wiped out. The general rule for the relationship between nose radius and feed rate is
that the nose radius should be three or more times the feed rate. However, it should
be remembered that too large a nose radius may induce chatter because a large nose
radius thins the chip and increases radial pressures (see Fig. 4.40c).

Feed Feed

Feed

Chip thinned
at this point

(a) (b) (c)

Fig. 4.40 Relationships between feed and nose radius

Closely related to the nose radius and its effect on surface roughness is the end
cutting-edge angle and the point of tangency between the nose radius and the end
cutting-edge angle. The finishing point of a single point cutting tool is actually the high
Design of Cutting Tools 253

point of the nose radius just prior to its blending into the end cutting-edge angle. The
point of tangency between the nose radius and the cutting-edge angle determines the
height of the feed marks on the surface. Thus, an increase in the end cutting-edge
angle increases the height of the feed marks and vice versa. Sometimes a small flat is
ground parallel to the work axis and ahead of the end cutting-edge angle. This serves
to further wipe out feed marks. Care must be taken to ensure that the tool is set in
such a manner that the flat will be parallel to the work axis.
The effect of the built-up edge on surface finish has already been discussed. Anything
that can be done to prevent or stabilise a built-up edge will generally improve sur-
face quality. Since the built-up edge is characteristic of slower cutting speeds in the
high-speed steel cutting range, higher cutting speeds may be a help. This is why
machine operators improve surface quality increasing cutting speed on a finishing cut.
Higher cutting speeds help to eliminate or stabilise the built-up edge. The speed may
be increased on light finishing cuts without additional tool wear because a smaller
amount of material is being removed and therefore less heat is generated. Surface
finish will improve with increased cutting speed up to a certain point, depending upon
the work material, tool material, tool shape, and cutting conditions. Thereafter, little
further improvement in finish occurs with increased cutting speeds.
The selection of cutting-tool materials may have an effect on the built-up edge. For
example, the improper selection of a carbide grade may result in the work material
welding to the tool tip and causing an increase in the built-up edge. Soft ductile mate-
rials will flow more freely over a finely finished surface. The same is true if a cutting
lubricant is applied to the tool face. This is particularly true in the speed range of high-
speed steel. Coolants or cutting lubricants may be detrimental if applied to carbides
and oxides because these tools are sensitive to sudden temperature changes and
fractured cutting inserts may result.

Chatter and vibrations A machine tool is usually constructed of localised masses,


each of which contain moving parts. Because of the wide range of feed and speed
combinations, the moving parts do not always move at the same velocity. The vari-
ous members of the machine-tool structure are all subject to loading in different ways
because of changes in work and cutter diameter, direction of feed, and varying depths
of cut. The result is a complete vibratory system having a very complex dynamic
behaviour. This vibratory system is of great concern to the machine-tool designer, as
vibrations are often detrimental to the workpiece, cutting tool, and the machine tool
itself.
The tool designer is interested in vibrations that affect the metal-cutting process. Such
vibrations are usually of two types: forced and self-excited. Forced vibrations are
those which occur under the action of rhythmically varying force on the cutting tool,
such as the rpm of an unbalanced pulley, spindle, workpiece, or grinding wheel. The
cutting tool or workpiece is actually pushed bodily by the varying force. A bad place
on a V belt could cause a forced vibration. Other causes could be unbalanced rotors
of electric motors or even the pulsing of hydraulic motors.
The frequency of a forced vibration corresponds to the frequency of the vibrating
agent and may be quite different from the natural frequency of the vibrating members.
Vibrations of this type usually cause problems for finishing operations such as grind-
ing, finish boring, and finish turning. They can be reduced by eliminating the disturbing
254 Tool Design

agent. One way of determining the frequency f is to measure the distance L between
chatter marks and divide this quantity into the speed v of the cutter in m/s. Then,
v
f = __
L
where f = frequency, cps
v = speed of cutter or work, m/s
L = measured distance between successively produced chatter marks
The frequency can also be measured by use of vibration-detection equipment. If the
frequency corresponds to the product of the number of teeth on a gear and its rota-
tional speed, the vibration is forced and caused by the contacts between the gear
teeth. If it corresponds to the rotational speed of a pulley or spindle, the rotating mem-
ber should be checked for balance.
Self-excited vibrations are resonant vibrations in which the cutter and workpiece
move with respect to each other at a frequency determined by the natural frequency
of the machine. This type of vibration is generally referred to as chatter and occurs
because of unstable equilibrium of the potential vibrating member (a boring bar, for
example). Thus, when force conditions acting upon a cutting tool cause it to vibrate
at a frequency near the natural frequency of the machine, a resonant condition is set
up in which a minimum excitation will produce a maximum amplitude. The cutting tool
will vibrate at the same frequency as the natural vibration within the machine but the
amplitude will “whip up” to extremely high levels. Once the condition is established,
the interaction between the cutter and workpiece will sustain the vibration. The con-
stant pounding will drastically reduce cutter life, impair surface finish, and in case of
high amplitude may cause damage to the workpiece and the machine tool itself.
The potential vibrating member may be excited by a forced vibration. For example,
the formation of a metal chip where the recurring fracture of the metal at the shear
plane produces rhythmic variations in force may cause the cutting tool or toolholder
to vibrate. This is a foced vibration in that the tool is made to deflect rhythmically.
However, if the frequency of deflection is near a natural frequency of the machine,
resonance will occur and the resulting chatter would be a self-excited vibration.
There are many factors involved in setting up the right conditions to produce chatter.
In fact, almost all the components of a machine tool and tool system may be involved.
The design of the machine tool, the type of workpiece material, the type of cutting
operation, the type and design of fixturing, the geometry of cutting tools, the selection
of feeds, speed, and depths of cuts, all go to make the conditions that produce chatter.
It is therefore difficult to pinpoint one main cause of chatter, as different materials may
react differently under the same cutting conditions. However, it is generally agreed
that rigidity of the machine tool and cutting system has a great deal to do with chatter
proneness. The tool designer cannot change the rigidity of the machine tool unless he
changes machines, but he is responsible for rigidity of cutting tools and workholding
devices. This usually boils down to tool overhang and cross-sectional area.
The tool designer should be aware of the physical laws having to do with relationships
between length and cross-sectional area. For example, the stiffness of a bar is pro-
portional to the fourth power of the diameter. Thus a 50 mm diameter milling-machine
arbour is 16 times as still as a 25 mm bar of the same length (504/254 = 16). A milling-
Design of Cutting Tools 255

machine arbour 30 mm in diameter would be over twice as stiff as a 25 mm arbour


(304/254 = 1.93). This example shows that the tool designer should use the largest
diameter possible consistant with the limitations of the workpiece and cutter.
Bar stiffness is also inversely proportional to the third power of the unsupported
length, as shown by the formula

F × L3
D = ______
3E × I
where D = deflection, mm
F = force, N
L = length of overhang, mm
E = modulus of elasticity of material
I = moment of inertia, base on shape and cross section of beam
Since L is cubed in the formula, the force required to deflect the tool will decrease by
the cube of its length.
The formula also shows that changing the tool material to a better grade of steel or
hardening to a higher hardness does not improve the stiffness of the tool since all
steels have practically the same modulus of elasticity and the addition of alloys or
heat treatment does not change the modulus value. The modulus of elasticity of most
steels is around 210 GPa. Thus, for a given tool shank, the load, length and depth
determine rigidity, and if they cannot be changed, the only quantity left is the modulus
of elasticity. A material with a higher modulus of elasticity should be selected. A good
example is tungsten carbide, which has a modulus of elasticity of approximately 650
GPa. It takes over three times more force to bend a carbide bar than that required
to bend heat-treated tool steel. The main advantage of heat treating a steel bar is to
prevent wear and physical damage to the bar during use.
When turning long slender bars or using boring bars of small diameter with excessive
overhang, it may be necessary to use a vibration damping technique. This technique
may be roughly compared to that of laying a finger on a vibrating violin string. In fact,
a machinist may press a small block of wood or lead weight against a rotating long
slender shaft or boring bar and reduce vibration in certain instances.
A commercial vibration damper available from Kennametal, Inc. is known as the
Kennametal Roller De Vibrator and is shown in Fig. 4.41. It consists of a wheel that
rolls on the outside diameter of the workpiece. The wheel is supported by a hinge
yoke and contacts the work just behind the cutting tool. The wheel is held on the work
by gravity.
The wheel contains inertia disks which are free to float within the wheel housing.
When the rotating workpiece begins to vibrate, the inertia disks make random impacts
against the inside of the wheel housing. These impacts, in turn, are transmitted back
to the workpiece through the contact wheel and set up a counterforce to damp the
vibration.
256 Tool Design

Fig. 4.41 Using the Kennametal Roller DeVibrator (Kennmetal, Inc.)

Several unusual innovations have been applied in the damping of vibrations when
using boring bars. One method frequently used is to hollow out the end of the bar and
fill it with a high-density material such as lead shot or mercury. A specific application
of this type is recommended by the Metallurgical Products Department of General
Electric Company. The damper consists of a piece of high-tungsten alloy (Hevimet)
that “floats” in a cavity located as near the tree end of the bar as possible. The cavity
and slug are accurately finished so that the slug will be supported by an air film. Inertia
of the slug damps out the vibration of the bar.
When designing a bar of this type, the damper diameter should always be as large as
the diameter of the bar will permit. For example, a 25 mm diameter boring bar should
have a damper 19 mm in diameter. The ratio of the diameter to length should be 1:2.
In the above example, the diameter of the damper is 19 mm, and so the length should
be 38 mm.
A 0.08 mm “float” tolerance should be allowed between the ground OD of the damper
and the ID of the hole in the boring bar. A 0.25 mm “float” tolerance should also be
provided by making the hole in the boring 0.25 mm deeper than the length of the
damper insert.
The easiest and most effective way to position the damper insert in the boring bar is
to drill and ream a hole of the required depth and diameter in the end of the bar, insert
the damper slug, and weld or braze a cap over the end (see Fig. 4.42). For use when
boring a blind hole, the damper insert should be positioned behind the tool, as shown
in Fig. 4.43. This can be accomplished through a cartridge-type arrangement. The
cartridge containing the tool is attached to the end of the boring bar, which serves as
a cap for the damper insert.

Tool Welded cap

Hevimet insert

Fig. 4.42 Vibration damping of a boring bar (Mettallurgical Products Department, General
Electric Company)
Design of Cutting Tools 257

Hevimet insert Tool

Weld or thread

Fig. 4.43 Vibration damping of a boring bar for blind holes (Mettalurgical Products Depart-
ment: General Electric Company)

A commercial device for damping vibration is used by the Austin Tool Company in
their Vibra/damp tools (see Fig. 4.44). The rear half of the bar shown in Fig. 4.44 is
of tubular tungsten carbide, while the steel front half incorporates a standard Vibra/
damp vibration damper. This damper makes use of coulomb, or friction, damping, as
shown in Fig. 4.45. A large-mass element located in the working end of the tool is
spring-loaded against a suitable bearing surface and supported concentrically within

Fig. 4.44 A vibration-damped composite boring bar for high-production


boring of difficult materials (Austin Tool Company)

a tapered cavity. Annular clearance is provided within the confines of the cavity, the
sole purpose of which is to prevent the mass from contacting the internal chamber
walls. A cavity in the large mass which is also spring loaded. The arrangement of
springs produces a larger contact pressure on the area labelled B than on the area
labelled C. The purpose of this is to provide for movement of both elements during
severe vibration, such as during rough cuts and also to allow limited movement of the
large mass only during light finishing cuts where the vibration is small. It is important
to remember that there must always be relative motion between the damping mass
elements and the bearing surface in order to generate coulomb friction, which pro-
vides the resisting force to vibration or chatter. This force, though small, operates at
such high frequencies that its damping effect is similar to that of laying a finger on a
vibrating violin string.
258 Tool Design

Fig. 4.45 Coulomb, or friction, damping system in a boring bar. Shaded areas indicate
relative motion with respect to bearing plate A: (a) Roughing cuts (b) Finishing
cuts (Austin Tool Company)

Another boring bar using the damping principle available from Kennametal, Inc. is
called the Kennametal De Vibrator-K-Bar and is shown in Fig. 4.46. The design prin-
ciple is similar to the Kennametal Roller De Vibrator. The shank of the bar is made of
a special grade of tungsten carbide almost three times as stiff as steel and has 30 per
cent greater transverse rupture strength and greater impact resistance than conven-
tional tungsten carbides.

Fig. 4.46 Using a Kennametal De Vibrator-K-Bar (Kennametal, Inc.)

The dynamic damping of vibration is accomplished by a number of inertia disks, as


shown in Fig. 4.47. A drawbar runs through the disks and through the full length of the
shank, holding the cutting head firmly to the end of the shank. The boring bar is shown
at an idealised starting point of vibration and is moving upward with all the disks rest-
ing at the bottom of the cavity. When the bar reaches the top of its cycle and starts
down the inertia of the disks carries them on upward until they hit the top of the cavity.
The impact of individual disks occurs with random timing. The light spring at the end
of the disks contributes to the random motion by causing a slight amount of friction
between the disks. Most of the impacts are opposed to the vibrational movement and
hence set up a counterforce that brings harmonic vibration to a stop.
Design of Cutting Tools 259

Fig. 4.47 Construction of the Kennametal De Vibrator-K-Bar (Kennametal, Inc.)

A boring tool manufactured by Vernon Devices, Inc. uses an entirely different prin-
ciple to prevent vibration (see Fig. 4.48). It is known as the Vernon Bore-Bit. This tool
is designed to support a boring bar at the cutting
edge and is reported to be six times as rigid as an
unsupported boring of similar proportions. It can
be used on machine tools with as little as 746 W
and is infinitely adjustable from 45 to 64 mm It will
bore within 0.05 mm of setting and is guaranteed
to provide 0.025 mm five inscribed radius at full
capacity. Out-of-roundness and angularity are
corrected in predrilled or cored holes. A built-in
coolant system forces coolant along the cutting
edges for maximum tool life. Its basic construc-
tion is shown in Fig. 4.49, and Fig. 4.50 shows the
Bore-Bit under actual cutting conditions.
Figure 4.51 shows how the Bore-Bit works. At a,
the trueing bit, independent of support, is about
to enter the starting hole in the workpiece. At b, Fig. 4.48 The Vernon Bore-Bit
the trueing bit has corrected initial misalignment (Vernon Devices, Inc.)
and out-of-roundness of the predrilled hole. The
nose cone (affixed so that it can slide) contacts the perimeter of the hole; coolant flow
begins when the nose cone engages the workpiece.
The boring bits engage the workpiece as the nose cone receives pressure from the
internal spring at c. The boring cutters now receive full support. Coolant flows to the
boring bits.
At d, the Bore-Bit generates the desired hole in a single pass. Note the three distinct
diameters: the drilled starting hole, the trueing hole, and the bored hole. The bore is
completed at e. All contact is automatically shut off.
The commercial devibrators are quite effective and should be used where vibration
and chatter occur. However, they are expensive and should not be used when stan-
dard boring bars will perform satisfactorily. Special consideration should be given to
the length-to-diameter ration of standard boring bars when selecting them for a partic-
ular job. Figure 4.52 shows the importance in proper length-to-diameter ratio of boring
bars. Seven 64 mm diameter boring bars are shown in suitable lengths so that their
length-to-diameter ratio is 1:1. 2, 3:1, etc. The figure at each tool point indicates the
force in pounds required to deflect the boring bar 0.025 mm. These figures are com-
puted for a heat-treated tool-steel boring bar, assuming a modulus of elasticity of 140
260 Tool Design

Coolant channel
Nose cone (for boring bits)
(chrome moly heat-treated)
Ports directing coolant Wrist pin
to clean and lubricate
HSS boring blades
bearings

Solid tungsten carbide HSS trueing bits


roller bearings
Trueing bit holder

Fig. 4.49 Construction of the Vernon Bore-Bit (Vernon Devices, Inc.)

GPa. Notice that a force of 49130 N is required to deflect the 64 mm bar (1:1 length-
to-diameter ratio) 0.025 mm. Further note that only 765 N deflects the 254 mm (4:1
length-to-diameter ratio) the same amount and that a force of only 32 lb is required
when the bar is 445 mm long with a 7:1 length-to-diameter ratio. It thus follows that
the boring bar with the 1:1 ratio is 49130/765 times as rigid as the bar with the 4:1
length-to-diameter ratio and 49130/142 times as rigid as the bar with the 7:1 ratio.
Thus, when specifying bar tooling, it is better to select bars whose length-to-diameter
ratio is favourable in maintaining the rigidity built into the machine-tool spindle.
Design of Cutting Tools 261

Fig. 4.50 Vernon Bore-Bit under cutting conditions (Vernon Devices, Inc.)

Stub boring bars for general-purpose use


on horizontal-boring machines and jig-
boring machines are tremendously more
rigid than the offset type of head. The
dramatic increases in rigidity can eas-
ily be seen in Fig. 4.53, which shows a (a)
stub bar at the left. Stub boring bars take
advantage of the inherent rigidity of the
machine tool itself to provide rigidity at
the cutting edge.
(b)
To summarise, chatter is generally due to
the lack of rigidity. The hammering action
against the cutting edges may cause pre-
mature tool failure and will certainly cause
a poor surface finish. If chatter does
occur, several elements may be consid- (c)
ered to help eliminate it. Looseness in
the machine due to wear may contribute
to chatter tendencies. The machine itself
may not be rigid enough for the prevention
of chatter. The rigidity of the work-holding (d)
fixture may be lacking, or the clamping
devices on the fixture may not be positive.
Tooling may lack rigidity. There should be
minimum overhang, and the shank diam-
eter should be as large as possible and
the length as short as possible.
(e)
In some cases, the tendency to chatter
Fig. 4.51 Cutting action of the Vernon
can be reduced by decreasing the cutting
Bore-Bit (Vernon Devices, Inc.)
speed or width of cut. Increasing the feed
may reduce chatter tendency in some cases, and reducing the cutter clearance may
have a similar effect. The use of a helical-flute cutter in place of one with straight flutes
may go a long way toward eliminating chatter.
262 Tool Design

Fig. 4.52 Importance of proper length-to-diameter ratios of boring bars. Figures represent force in
Newtons required to deflect boring bar 25 mm at lengths shown; length-to-diameter ratio.
Figures at tool points indicate the force in Newton required to deflect each bar
25 mm. (Devlieg Microbore)

Fig. 4.53 Using a stub boring bar (left) greatly increases rigidity (Devlieg Microbore)

Cutting fluids Cutting fluids, often referred to as coolants or cutting lubricants, gen-
erally have a two-fold function: (i) They lubricate and serve to reduce friction between
the cutting tool and workpiece, and (ii) They cool and carry off heat generated at the
cutting edge. Some cutting fluids have greater lubricating qualities while others have
greater cooling qualities. This is probably why some cutting fluids are referred to as
coolants while others are referred to as lubricants.
It is not known when fluids were first used as an aid to metal cutting. Early machine
operators probably found that a somewhat better finish was secured by the occa-
sional application of ordinary machine oil and that a still better finish was obtained
Design of Cutting Tools 263

when using lard oil. It is also probable that early operators used plain water for cooling
tools when they were taking heavy cuts.
The first formal research was conducted by FW Taylor in 1883. He showed that a
heavy stream of water applied to the cutting edge would allow cutting speeds to be
increased by 30 to 40 per cent. Later, animal oils, ordinary machine oils, and water
solutions of soda or soap were used as cutting fluids. The animal oils, often referred to
as fatty oils (lard oil, sperm oil), had the disadvantage of turning rancid and producing
foul odours under conditions of shop use. To combat this tendency, animal oils were
mixed with mineral oils to produce fatty mineral oils that worked very well for opera-
tions where service was not especially severe. Emulsions of oil (soluble oil) replace
soda and soap in water where maximum cooling was necessary.
In recent times, the need for higher production and the introduction of new materials
have created a demand for improved cutting fluids. Additives have been developed
that give specific properties to cutting fluids. Sulphur and chlorine are the common
additives in mineral oils. The oil emulsions in water-soluble cutting fluids have been
replaced with chemicals that contain no mineral oil, and they either dissolve in water
or form colloidal solutions, where the particles are held in suspension. The result
is that today there are many blends and compounds to choose from. The choice
between “goat’s milk” (soluble oil in water) and “thread-cutting oil” is no longer appli-
cable. Today’s cutting fluids are tailor-made to give specific properties to cutting fluids,
and the tool or manufacturing engineer must be aware of available fluids, fluid types,
fluid function, and selection and methods of application.
The basic functions of metal-cutting fluids are to lubricate and cool. Secondary ben-
efits of a metal-cutting fluid include washing away chips, protecting the work against
rusting, lubricating machine ways, and controlling dirt and dust, as in the case of
machining cast iron.
Limitations restrict the use of some cutting fluids. Petroleum-based fluids may present
a fire hazard. Many fluids are susceptible to bacterial attack, especially when they are
allowed to stand in machine reservoirs for periods of time without use. Certain fluids
may cause corrosion of workpiece material, while others may attack the skin of the
machine operator.
The broad general classification of cutting fluids is water-based and oil-based. All
cutting fluids provide a cooling and lubricating action, the degree of cooling and lubri-
cating depending upon the physical makeup of the fluid itself. To cool effectively, a
fluid must have a high specific heat, a high heat of vapourisation, and a high thermal
conductivity. Pure water is probably the best coolant. It has a specific heat of 1.0.
Hydrocarbon oil has a specific heat of 0.45, which means that an equal weight of
water compared to oil will have a resultant lower temperature for a given amount of
heat input. However, water promotes rust and has practically no lubricity at all.
The lubrication of two surfaces sliding relative to each other may be classified as
either hydrodynamic (fluid-film) or boundary lubrication. In hydrodynamic lubrication,
the surfaces are completely separated by a fluid film; the chemical nature of the fluid
is immaterial, only the viscosity of the fluid being important. This can be compared to
applying a film of oil to a surface plate and sliding a metal block across it. In boundary
lubrication, the surfaces are separated by a physically or chemically adsorbed film of
a few molecules in thickness. Adsorption is the adhesion of the fluid molecules to the
solid surfaces with which they are in contact. Here the chemical nature of the cutting
264 Tool Design

fluid is important, since the polarity of the fluid molecules determines the affinity with
which they are held on the surface. The chemical nature of these adsorbed films is
such that they are generally tough and resist removal from the surface, even under
high pressure and at high velocities of the rubbing surface.
Another form of boundary lubrication occurs when certain elements in the cutting
fluid react chemically with the solid surface to form a compound with lower shear
strength than the solid itself. This is often referred to as chemical or extreme-pressure
lubrication.
It is important that these basic lubrication principles be understood before studying
metal-cutting fluids. For example, it is known from previous study of chip formation that
the forces at the chip-tool interface may be quite high. These high pressures, along
with velocities and higher temperature, set up conditions that make hydrodynamic
lubrication impossible. Therefore, boundary lubrication, either of the adsorbed-film or
chemically formed film type, is important when considering metal-cutting fluids. Pure
petroleum oil is generally considered as a lubricant, but when modern cutting fluids
are concerned, it may be only a carrier for active chemicals that set up boundary-
lubrication conditions.
The dissipation of heat by conduction has already been mentioned. The lubrica-
tion qualities of a metal-cutting fluid will also lower tool temperature by reducing the
amount of heat generated. The bulk of heat produced in the formation of a metal chip
is produced in the area of shear, where heating is caused by deformation of the metal.
It is estimated that 60 per cent or more of the total heat is generated in this area; 30
per cent is produced by friction where the chip rubs across the face of the tool; and
the remaining 10 per cent is produced at the flank of the tool.
When a cutting fluid with good lubricating ability is applied, friction between the chip
and tool is decreased. This allows the chip to slide up the face of the tool, and the
shear angle in turn is increased. The large shear angle produces a short shear path,
which deforms less metal. The end result is a longer, thinner chip and considerably
less heat.
The reduction of friction between the tool face and the chip will also help control the
built-up edge. The chip metal tends to slide over the tool face instead of adhering to
the cutting edge.
Cutting-fluid manufacturers have hundreds of cutting-fluids types under various trade
names in the market. However, most cutting fluids on the market fall in one of the
three classes. These are the petroleum or mineral-based cutting oils, the emulsifiable
(soluble) cutting oils, and the chemical or synthetic cutting fluids.
The petroleum or mineral-based cutting oils may be a “straight” cutting oil with very
few additives or a highly compounded blend specifically formulated for a special oper-
ation. Straight cutting oils without additives are used very little today. They are occa-
sionally used for light-duty machining of free-cutting metals, but for the most part act
as a carrying vehicle for the added chemicals. Broadly speaking, the petroleum-based
cutting-oil blends are as follows.
Fatty mineral oils Fatty mineral oils consist of one or more fatty oils blended into
straight mineral oils. Lard oil is usually the main fatty oil additive. They provide wet-
ting and oily properties not found in straight mineral oils. These blends are used
Design of Cutting Tools 265

extensively for automatic screw-machine work, where the service is not especially
severe and where high surface finish and precision are required. An additional advan-
tage is that they do not corrode either ferrous or nonferrous metals.
Sulphurised mineral oils In contrast to the fatty mineral oils, sulphurised mineral oils
are highly active and therefore will corrode or discoluor copper alloys. An “active” or
“inactive” oil refers to the chemical activity of the oil at elevated temperatures. In the
sulphurised mineral oils, the heat in the cutting zone releases some of the sulphur to
react chemically with the work surface to provide boundary lubrication when machin-
ing steels. The sulphur can react to form iron sulphide, which has a lower coefficient
of friction than the parent metals. Sulphurised mineral oils are best used in machining
gummy and highly ductile low-carbon and plain steels that are prone to form a built-up
edge on the cutting tool.
Sulphurised fatty mineral oils These blends are a compromise between straight fatty
mineral oils and sulphurised mineral oils. They are inactive at lower temperatures but
become active at temperature and pressure conditions of the chip-tool interface. For
this reason, they are nonstaining oils suitable for machining both ferrous and non-
ferrous metals on the same machines. They have excellent antiweld characteristics
along with high lubricity, characteristic of the fatty oils.
Sulphochlorinated mineral oils A chlorine additive is sometimes introduced into
cutting oils in combination with sulphur. Chlorine reacts at the chip-tool interface in
essentially the same manner as sulphur to provide boundary lubrication. Its main
advantage is that it is more reactive than sulphur and, by comparison, it becomes
effective at lower temperatures. A sulphochlorinated mineral oil is used whenever
antiweld and pressure-resisting properties must be effective over a wide temperature
range. They are especially suited for soft, draggy metals and chrome-nickel alloys.
The emulsifiable or soluble cutting oils are mineral oils that contain soap or other
chemicals to make them mix with water. Actually, the oil does not mix with water but
becomes a suspension of minute oil droplets in water. Since water has high specific
heat, high thermal conductivity, and high heat of vapourisation an, an emulsifiable
cutting fluid is one of the most effective cooling media known. However, since the
viscosity of the emulsion is nearly equal to that of water, its inherent film strength and
lubricating properties are not as good as most straight cutting oils. Emulsifiable oils
are therefore best suited for metal-cutting operations with high cutting speeds and low
cutting pressures, where considerable heat is generated. By varying the soluble-oil-
water mix ratio, the lubricating or cooling properties can also be varied. Leaner emul-
sions may be used in machining operations (grinding) that require fast heat removal,
while operations requiring high lubricating properties (milling) may call for a richer
mixture. Rust prevention is also increased by higher concentrations of soluble oil.
There are three types of emulsifiable cutting oils in use today:
Regular Mineral-Oil Emulsion The emulsifiable mineral oils are the most widely
used of the soluble oils. They consist of a straight mineral oil made emulsifiable by
the addition of soap or a soap-like material to make them “soluble” in water.
Fatty Emulsifiable Mineral Oils These oils are similar to the straight emulsifiable
mineral oils except that a fatty oil (lard or sperm oil) has been added to increase the
oiliness of the emulsion.
266 Tool Design

Extreme-Pressure Emulsifiable Oils These oils contain extreme-pressure additives,


such as sulphur and chlorine, to impart boundary-lubrication properties for machining
tough and gummy metal. They are often referred to as heavy-duty soluble oils and in
many instances have replaced mineral-oil-base cutting fluids. Their advantage is that
they cool better and do not smoke as much as oils under heavy cuts.
Chemical synthetic fluids do not contain mineral oil, and they either dissolve in water
to form a true solution or form colloidal solutions, where the particles are held in sus-
pension. They have excellent cooling qualities because they are mixed with water.
The chemical agents that go into the fluid include rust inhibitors, water softeners, mild
lubricants, germicides, wetting agents, and water conditioners. In general, they are
divided into two groups: (i) Those with wetting agents and lubricants and (ii) Those
without. Group 1 has excellent lubricating properties and good corrosion or rust
resistance. It is suitable for machining operations requiring high lubricating qualities
such as reaming, milling, sawing, tapping, broaching, etc. The addition of extreme-
pressure additives (chlorine, sulphur, phosphorus) along with the wetting agent may
impart boundary lubrication properties for tough machining jobs.
Group 2, without the wetting agent, is best used as a coolant and is well adapted to
grinding operations. It is a good rust inhibitor but lacks lubricity. It has a tendency to
leave a hard or crystalline deposit when the water evaporates and may interfere with
the action of the machine. It may also attack the paint used on the machine tool.
The important factors to consider when selecting a cutting fluid are the nature of the
cutting operation and the materials being machined. For example, a broaching or
threading operation where severe frictional problems are expected will require a fluid
having high lubricity and antiwelding properties. In this case, an oil-type fluid contain-
ing extreme-pressure additives would be the best choice. On the other hand, a sur-
face grinding operation, where high cooling is necessary, would call for a water-base
cutting fluid. An oil emulsion or synthetic grinding fluid would be the logical choice. An
emulsion with extreme-pressure additives may be used where both cooling and anti-
welding qualities are required. When a particular machine is being used to machine
both ferrous and nonferrous materials, a cutting fluid which would not stain either
material may be the proper selection. When fogging and smoking are objectionable,
a water-base fluid may be a better choice. The cost of cutting fluid may also influence
the selection. It is uneconomical to pay extra for an oil-base cutting fluid containing
expensive additives when a less expensive emulsion will do the job satisfactorily.
Table 4.14 gives general recommendations for the selection of cutting fluids.
The care taken in the selection of a cutting fluid is a wasted effort if the fluid is not
applied correctly. All too often, the machine operator applies the cutting fluid in insuffi-
cient quantities, intermittently, or to an area where it will do no good. When machining
with carbide or oxide cutting tools, an insufficient or intermittent flow of coolant may
actually be detrimental to tool life. This is the result of alternate heating and cooling
of the cutting tool, which in turn subjects it to thermal shock. Minute cracks appear in
the surface of the tool insert, causing small pieces of the cutting edge to break away.
In this case, it would be better to machine dry.
Cutting fluids may be applied by hand, by flood, or by mist. Hand application is gener-
ally accomplished by use of a pump-type can or brush and is limited to slow-speed
applications where boundary lubrication is desired. A tapping operation is typical of
this method.
Table 4.14 Recommended cutting fluids
Metal to be machined
Machining operation Steel Cast iron Nickel Copper Brass Aluminium Magnesium
and alloys and bronze
Low-carbon High-carbon Stainless
and alloy

Broaching J J J D10 J B C C B
Threading H J J D10 J B C C B
Gear cutting E, K10 F J D20 ... ... B
Drilling and reaming E, D10 F, K10 J D20 E B B B B
Boring and turning D20 K10 K5 ... E D30 D30 D30 B
Automatic screw machining E E, H H, J ... H A B C B
Milling D20 D20 K5 D20 F D30 B D30 B
Thread rolling F F F ... ... A C B A
Sawing D30 D30 D30 D30 D30 D30 D30 D30 B
Thread grinding J J J ... ... ... C C C
Other grinding D50 D50 D50 D40 D50 D50 D50 D50 B
Honing J J J
Key
Inactive types Active types
A = mineral oil E = sulphurised mineral oil
B = mineral-fatty oil F = sulphurised mineral-fatty oil, light
C = mineral oil plus sulphurised fatty oil G = sulphurised mineral-fatty oil, heavy
D = soluble oil (number after designated grade recommendation H = sulforchlorinated mineral oil
indicates water dilution: D20 signifies a 1:20 oil (water mixture) J = Sulfochlorinated mineral-fatty oil
K = heavy-duty soluble oil (see D)
NOTE: Because of fire hazard, do not use emulsions for machining magnesium.
Design of Cutting Tools 267

SOURCE: ASTME, “Tool Engineers Handbook,” 2d ed., F W Wilson (ed.), McGraw-Hill, New York, 1959, by permission.
268 Tool Design

Flood application is the most common method. A coolant pump delivers the cutting
fluid from a storage reservoir to a nozzle which directs the stream over the cutting
zone. The fluid is then collected in the T slots of the machine table and chip pan and
returned to the storage reservoir.
The flow of coolant must be directed to the point of cutting action in sufficient quanti-
ties to do its intended job. The coolant pump should provide enough force to penetrate
between the chips being formed and the cutting edges in the case of a multiple-tooth
cutting tool. This will also aid in washing chips away from the cutting area, especially
when machining in pocketed areas.
A single coolant nozzle is not sufficient when machining with larger milling cutters.
Milling cutters of large diameter but narrow widths (side mills, staggered-tooth mills,
and large saws) require a coolant nozzle on each side to ensure a sufficient quantity
of fluid (see Fig. 4.54). Fan-shaped nozzles may be used for wide-slab milling cutters
and gang mills, as shown in Fig. 4.55. Large face-milling cutters may utilise a ring-
type nozzle that distributes cutting fluid around the entire periphery of the cutter (Fig.
4.56.)

Fig. 4.54 Use of two nozzles for flood application of large-diameter and
narrow-width cutters

A single flood coolant nozzle is not efficient when drilling relatively deep holes. The
flutes of the drill, designed to auger the chips from the hole, also remove coolant. The
operator must continually retract the drill to allow cutting fluid to enter the hole. This
practise may not be effective when the drill and workpiece lie in a horizontal plane.
The only effective method of application is to use oil-hole drills in which a passage
is provided through the body of the drill to allow cutting-fluid flow. The fluid is then
pumped under pressure through the drill to the cutting edge and is in turn forced out
of the hole through the drill flutes.
Grinding is another machining operation that cannot be efficiently lubricated by a
single coolant nozzle. This is due to the high speed of the grinding wheel, which actu-
ally “fans” the cutting fluid away from the cutting area. The rotating wheel acts as an
air pump discharging air out of its periphery.
A common method of getting fluid into the grinding zone is to pass it through the grind-
ing wheel. Since a grinding wheel is quite porous, water-base cutting fluids easily
pass through it. Cutting fluid may be introduced directly through a hollow spindle into
the grinding-wheel bore, in which case one must purchase grinding wheels without
bushings or painted bores.
Design of Cutting Tools 269

Workpiece Workpiece

Fig. 4.55 Use of two fan-shaped nozzles Fig. 4.56 A ring-type coolant nozzle for
for flood-coolant application face milling
on wide slab milling cutters

Another method is to introduce the cutting fluid through the side of the grinding wheel.
The spindle flanges are replaced with modified flanges that have a “slinger” groove
machined into them. A series of holes is drilled into the bottom of the groove to allow
coolant passage into the side of the wheel (see Fig. 4.57).
The cutting fluid must be filtered before reuse because small particles of swarf will
plug the openings in the grinding wheel. Grinding-machine manufactures recommend
a 3 mm filter. A great deal of fog is also present during use because cutting fluid is
discharged around the entire periphery of the wheel.
A second method of getting cutting fluid into the grinding zone is by the use of spe-
cially designed nozzles. Figure 4.58 shows an example. Here the cutting fluid is in
contact with the grinding wheel long enough to attain a high velocity and be forced
into the wheel face and grinding zone. The width of the nozzle should be approxi-
mately three-fourths the width of the grinding wheel.
270 Tool Design

Coolant
flow

Slinger ring

Grinding
wheel

Grinding
wheel

Fig. 4.57 Introducing cutting fluid through the side of the grinding wheel

Mist application of cutting fluids is accomplished by using a mist generator, which


basically consists of a container for cutting fluid with attached hose and nozzle. The
cutting fluid is atomised by the specially designed nozzle using compressed air from
the shop line. The nozzle directs the mist spray on the cutting edge of the tool. The
heat of the tool and the work is dispersed by the rapid movement of the expanding
compressed air and by evaporation of the
coolant. Water-base cutting fluids are gen-
erally used because of their high heat of
vapourisation.
Mist generators in common use are of
the aspirator or the pressure-fed type.
The most common and least expensive
is the aspirator type, which works by the
same principle as the common household
sprayer. A stream of air is blown over the
open end of a tube within the nozzle. The
air passing through the nozzle creates a
partial vacuum which syphons the coolant
from the container. There is no pressure
in the container at any time. The nozzle is
adjustable to regulate the airflow; thus the
amount and density of mist can be partially Fig. 4.58 Coolant nozzle designed
controlled. Since only a partial vacuum is to maintain longer contact
formed, the hose between the spray noz- between cutting fluid and
zle and the cutting-fluid container must be grinding wheel
Design of Cutting Tools 271

relatively short for successful operation. The


container is usually mounted on each individ-
ual machine relatively close to the cutting tool.
Figure 4.59 shows a typical aspirator-type mist
generator.
The pressure-fed type of mist generator uses
direct pressure rather than a vacuum to feed
the cutting fluid into the airstream. The cutting-
fluid container may be a considerable distance
from the machine, and individual valves may
be installed at each machine on the air and
cutting-fluid supply lines. Hand valves may be Fig. 4.59 A typical aspirator-type
used, but no production machines where an mist generator (Trico
intermittent mist is required, it is more conve- Fuse Mfg. Co.)
nient to install electrically controlled solenoid
valves. They may be connected to a lighting
circuit or directly across the motor starting
switch to start or stop the mist spray when the
motor is turned on or off. Figure 4.60 shows a
pressure-fed mist generator.
Mist cooling has been used on all types of met-
al-cutting operations, but it is generally used
on small-diameter cutting tools operating in the
higher speed ranges.
Typical examples are end-milling and drill-
ing operations. It is also particularly useful
on machines without built-in cooling systems
because of the low cost and ease of installa- Fig. 4.60 A pressure-fed mist
tion. The cutting fluid is not reused and there- generator (Bijur Lubri-
fore does not require a recirculation system. cating Corporation)
Care must be taken to prevent creating a
health hazard when using mist coolants, as the mist particles may be inhaled by the
operator. Proper ventilation should be provided, especially when a large number of
generators are in operation in a confined area.

4.6 METAL-CUTTING TOOLS


The preceding sections have been concerned with the general aspects and consider-
ations of the metal-cutting process. Single-point cutting tools have served to illustrate
metal-cutting principles, other types of cutting tools being cited at appropriate places.
The following sections are concerned with specific metal-cutting tools with special
emphasis on standardised items that can be purchased commercially.
Consider first the various methods used to work metals. In general, these methods
may be divided into two classes: cutting tools and forming tools. Cutting tools remove
metal in the form of chips; forming tools either deform metals by making them flow into
new shapes or by shearing them into new shapes. Cutting tools are the subject of this
chapter, and forming tools are discussed in other chapters.
272 Tool Design

Cutting tools, or tools that remove metal in the form of chips, may be divided into five
basic groups according to machining operations. They may be classed as turning,
drilling, milling, shaping or planing, and grinding tools. Chip formation is the same in
all operations, but the method of holding the work is different. Each operation pres-
ents its own problems with regard to the cutting tools, but in many cases will be the
same because of similarity in machining operations. The following sections will dis-
cuss the common cutting tools used for the basic machining operations.

4.7 SINGLE-POINT CUTTING TOOLS


Single-point metal-cutting tools are normally used when performing turning, boring,
shaping, or planing operations. In the past, they were shaped from a solid bar of
cutting-tool material and were mounted directly onto the machine tool during the met-
al-cutting operation. The introduction of expensive cutting-tool materials has neces-
sitated that the body of the tool be made of a less costly material with a tip of cutting
material attached at the cutting point. The cutting-tool materials commonly used are
high-speed steel, cast alloy, cemented carbide, and cemented oxide. The tip or cut-
ting insert is held in the proper position mechanically or permanently soldered or
brazed in position. It may be presharpened or be an unground blank that is sharpened
after installation in the cutter body.

Side End cutting-


clearance edge angle
Heel Side rake
Face

Nose radius Side cutting-


edge angle

Side relief
Back rake

Shank
Flank

End
clearance
End relief

Fig. 4.61 General nomenclature for a single-point cutting tool

Figure 4.61 shows the general nomenclature for a single-point cutting tool. Single-
point tool elements and tool angles are defined as follows:

Shank Main body of the tool.

Size Determined by the width of the shank, the height of the shank, and the total
overall length.

Flank The surface or surfaces below and adjacent to the cutting edge.
Design of Cutting Tools 273

Heel The intersection of the flank and the base of the tool.

Face The surface on which the chip impinges.

Cutting edge The portion of the face edge that separates the chip from the work-
piece. The total cutting edge consists of the side cutting edge, the nose, and the end
cutting edge.

Nose The intersection of the side cutting edge and end cutting edge.

Side cutting-edge angle The angle between the side cutting edge and the side of
the tool shank. It is often referred to as the lead angle.

End cutting-edge angle The angle between the end cutting edge and a line perpen-
dicular to the shank of the tool.

Side relief angle The angle between the portion of the side flank immediately below
the side cutting edge and a line perpendicular to the base of the tool, measured at
right angles to the side flank. It is the angle that prevents interference as the tool
enters the work material.

End relief angle The angle between the portion of the end flank immediately below
the end cutting edge and a line perpendicular to the base of the tool, measured at
right angles to the end flank. It is the angle that allows the tool to cut without rubbing
on the workpiece.

Side clearance angle A secondary angle directly below the side relief angle, mea-
sured with the same reference.

End clearance angle A secondary angle directly below the end relief angle, mea-
sured with the same reference.

Back-rake angle The angle between the face of the tool and a line parallel with the
base of the tool, measured in a perpendicular plane through the side cutting edge.
It is the angle which measures the slope of the face of the tool from the nose toward
the rear. If the slope is downward toward the nose, it is negative back rake; and if the
slope is downward from the nose, it is positive back rake. The back-rake angle is zero
if there is no slope.

Side-rake angle The angle between the face of the tool and a line parallel with the
base of the tool, measured in a plane perpendicular to the base and the side cutting
edge. It is the angle that measures the slope of the tool face from the cutting edge. If
the slope is toward the cutting edge, it is negative side rake; and if the slope is away
from the cutting edge, it is positive side rake. If there is no slope the side-rake angle
is zero.
This single-point tool nomenclature is generally standardised. The tool angles are
taken with reference to the cutting edge and are therefore normal to the cutting edge.
They are of particular importance to the tool designer and machinist as they are the
ones specified when designing or grinding a single-point tool. A convenient way
to specify tool angles is by use of a standardised abbreviated system called tool
274 Tool Design

signature (sometimes referred to as tool character). Tool signature also describes


how the tool is positioned in relation to the workpiece. In other words, a tool signature
indicates the angles that a tool utilises during the cut. This tends to cause confusion
in that the angles that the tool utilises may not be the angles on the tool shank. For
example, a 45° side cutting-edge angle boring tool held in a 45° boring bar cuts with a
0° side cutting-edge angle; therefore, in the tool signature the side cutting-edge angle
would be listed as 0°. For the most part, however, a tool signature specifies the effec-
tive angles of the tool normal to the cutting edge. This will always be true as long as
the tool shank is mounted at right angles to the workpiece axis.
A typical tool signature will always appear in a definite order, as shown in Fig. 4.62.
The order used is back rate, side rake, end relief, end clearance, end cutting edge,
side cutting edge, and nose radius. End clearance and side clearance are omitted
unless clearance is used below the relief angles. Parentheses are placed around
these values when they are included in the tool signature.

Fig. 4.62 Typical tool signature

The basic angles on a single-point cutting tool vary in magnitude and direction. The
selection of the various combinations of tool angles depends upon many features
such as cutting-tool material, work material, workpiece set-up, depth of cut, rigidity of
machine tool, etc. The beginning tool designer may rely to a degree upon handbook
recommendations for basic angles, but the handbook cannot judge such variables
as workpiece and machine-tool rigidity. It is therefore necessary that the designer
has a general understanding of the effects basic tool angles have on single-point tool
performance.
The Side Cutting-Edge Angle (SCEA) or lead angle may vary from 0 to 90°. One
effect of increasing the SCEA is that the full length of the cutting edge is not in contact
with the workpiece when the tool enters the cut. The tool takes a small shock load and
gradually reaches the full depth of cut without undue impact. With a 0° lead angle, the
full length of the cutting edge is in contact with the workpiece at once and produces
a strong initial shock. By the same token, the tool forces are gradually decreased as
Design of Cutting Tools 275

the tool leaves the cut. This prevents chipping castings or “pushing off a ring” in the
case of steel alloys. Increasing the SCEA also increases the forces of right angles
to the work and may help hold the cross slide tight against the feed screw to aid in
preventing vibration in older machine tools. However, the same force tends to push
the workpiece away from the cutting tool. Thin-wall tubing or long slender workpieces
may be more satisfactorily machined with decreased SCEA because it reduces the
force at right angles to the work.
Increasing the SCEA decreases the thickness of the chip and increases its width.
For a given depth of cut, the material will be spread over a greater cutting surface,
as shown in Fig. 4.63. A SCEA of about 15° will reduce the power requirements and
will allow increased feeds. However, if the angle becomes too large, a point may be
reached where chatter will result because of excessive tool-contact area. Also, the
chip becomes thinner, and the downward cutting pressure moves outward toward
the cutting edge and increases the possibility of chipping. This is particularly true
in machining difficult metals, but the effect may be partially offset by increasing the
feed.

Fig. 4.63 Relative chip thickness as determined by side cutting-edge angle

The shape of the workpiece will often determine the SCEA. A 0° SCEA will produce a
90° shoulder. A negative SCEA angle of 2 to 5° is sometimes used to trace out a long
90° shoulder. However, unless the workpiece shape determines the SCEA, an angle
of 15 and 30° is an excellent choice for general machining. Standard tools are readily
available in these angles.
The End Cutting-Edge Angle (ECEA) prevents rubbing between the end of the tool
and the workpiece. Common ECEAs are 8 and 15°, although they may vary from 4
to 30°. Angles of 4° or less tend to cause vibration because of excessive tool contact
with the work. A rigid set-up is required because small ECEAs do not allow the tool
to penetrate the workpiece readily (excessive pressure normal to the workpiece are
produced). However, the resulting blocky tool point is much stronger. Greater angles
sometimes work quite well with frail workpiece set-ups where the forces on the tool
are not excessive in the first place. ECEAs of 30° or more are quite common for trac-
ing tools where the cross and longitudinal feed are operated simultaneously.
The End Relief Angle (ERA) and Side Relief Angle (SRA) prevent rubbing below the
cutting edge. Relief angles range from 5 to 15° for general turning. Small relief angles
276 Tool Design

give maximum support below the cutting edge and are necessary when machining
hard and strong materials. Increased relief angles reduce the forces required to pen-
etrate the workpiece material and cut more efficiently; they are commonly used for
materials with lower tensile strength. Too little relief can be as detrimental to tool life
as too much. If the flank of the tool rubs the workpiece, high forces and friction will
cause poor finish and overheating, and rapid wear will result. Too much relief weak-
ness the cutting edge, and failure may take place by chipping or deformation. The
small cutting-edge angle will also be affected to a greater extent by heat since there
is less mass to absorb and conduct the heat away from the cutting edge.
Clearance angles (sometimes called secondary angles) may be provided below relief
angles to prevent interference between the heel of the tool and the work, as in boring
or circular tools. In carbide tools, clearance is provided below the relief angle on the
carbide insert to prevent contact between the steel shank and the diamond grinding
wheel used to finish the carbide insert.
The rake angles of a single-point tool tend to determine the direction of chip flow
across the face of the tool. A positive back-rake angle tends to move the chip away
from the machined workpiece surface, while a negative back rate tends to move the
chip toward the machined workpiece surface. Cutting efficiency is best with positive
rake angles since the tool penetrates the workpiece easier and tends to shear the
metal off rather than compress and push it off. However, positive rake angles also
result in a more fragile cutting edge. The blocky cutting edge produced by negative
rake angles is stronger and is commonly used on materials where increased tool
strength is needed. Negative rake angles are most effective when used with tool
materials capable of machining at high speeds because rake angles have less effect
on tool pressures in the higher speed ranges. Negative rake angles are a necessity
when taking interrupted cuts with carbide cutting tools. In taking an interrupted cut the
shock load occurs back of the cutting edge, where there is more strength.
Back rake has less influence on the cutting action than side rake. Standard brazed
carbide tools commonly use 0° back rate. The side rake is generally varied from 0 to
15° to provide the necessary shearing action.
General rules for rake angles are as follows:
1. Use positive rake angles when machining low-strength ferrous and nonferrous
metals.
2. Use positive rake angles when machining work-hardening materials.
3. Use positive rake angles when using small machines with low power.
4. Use positive rake angles for turning long shafts of small diameters.
5. Use positive rake angles when the setup and workpiece lack rigidity and
strength.
6. Use positive rake angles when it is necessary to machine below recommended
cutting speeds.
7. Use negative rake angles for machining high-strength alloys, providing the
machine tool has the necessary horsepower.
8. Use negative rake for interrupted cuts and heavy feed rates.
9. Use negative rake for rigid setups in the carbide and oxide cutting-speed
ranges.
General recommendations for single-point turning-tool geometry are given in Table
4.15.
Table 4.15 General recommendations for geometry of single-point turning tools
All values in degrees

Material BHN High-speed-steel and cast-alloy tools Carbide tools


Brazed Throwaway
Back Side End Side ECEA Back Side Back Side End Side ECEA
rake rake relief relief rake rake rake rake relief relief

Gray or flake graphite cast iron 140 5 10 6 6 6 0 6 5 neg 5 neg 5 5 5


Nodular or ductile cast iron 180 3 8 5 5 5 0 6 5 neg 5 neg 5 5 5
Malleable cast iron 220 0 5 5 5 5 5 negg 5 neg 5 neg 5 neg 5 5 5
Free-machining plain carbon steel, 180 10 12 8 8 5 0 6 5 neg 5 neg 5 5 5
plain carbon steels
Free-machining alloy steels 250 8 10 6 6 5 0 6 5 neg 5 neg 5 5 5
Alloy steels, cast steels 350 0 8 5 5 5 0 6 5 neg 5 neg 5 5 5
Hot-work die steels, tool steels 500 0 5 5 5 5 5 neg 5 neg 5 neg 5 neg 5 5 5
Ferritic stainless steel 180 5 8 6 6 6 0 6 0 5 5 5 5
Austenitic stainless steel 200 5 6 6 6 6 0 6 0 5 5 5 5
Martensitic stainless steel 180 5 8 6 6 6 0 6 0 5 5 5 5
440 0 5 5 5 5 0 6 5 neg 5 neg 5 5 5
Precipitation-hardening stainless steel 220-320 0 5 5 5 5 0 6 0 5 5 5 5
Aluminium alloys 40*-110 20 15 12 10 5 3 15 0 5 8 8 5
Magnesium alloys 30*-80 20 15 12 10 5 3 15 0 5 8 8 5
Copper alloys 120-185 5 10 8 8 5 5 8 0 5 5 5 5
Titanium alloys 280-360 0 5 5 5 5 0 6 0 5 5 5 5
High-temperature alloys 260-320 5 6 5 5 5 0 6 0 5 5 5 5
*s500-kg load.
Design of Cutting Tools 277

SOURCE: ASTME, “Machining with Carbides and Oxides,” F. W. Wilson (ed.). McGraw-Hill, New York, 1962, by permission.
278 Tool Design

The importance of the nose radius is often overlooked when considering the single-
point cutting tool. The nose radius strengthens the tool point by thinning the chip
where it approaches the point of the tool and by spreading the chip over a larger area
of the point. It also produces a better finish because the toolmarks are not as deep as
those formed with a sharper tool. The nose radius is usually 0.79 to 1.59 mm. A larger
nose radius may be permissible if both tool shank and workpiece are rigid. A larger
nose radius is desirable during heavy depth of cut, heavy feeds, and interrupted cuts.
However, too large a nose radius will cause chatter because of the length of contact
between the cutting edge and one workpiece. A small nose radius is necessary if the
work cannot be held securely, if it is long and lacks rigidity, or if it is of thin-walled
construction.
General recommendations for nose radius are given in Table 4.16.

Table 4.16 Nose-Radius recommendations


Depth of cut, mm Nose radius, mm

3.18 0.79
4.77 – 9.53 1.19
11.12 – 19.07 1.59
20.65 – 31.78 2.38

The basic angles describe the shape of the cutting tool and are meaningful in tool
specification, tool maintenance, and manufacture. They are readily measured with
a protractor, and tool-grinding equipment is easily adjusted to produce each of the
basic tool angles. Each basic angle is considered one at a time when it is ground
or shaped into the tool. However, under actual cutting condition, each basic angle
combines with the others to form compound functional angles which substantially
determine the cutting performance. It is often difficult to visualise and define various
compound angles that determine chip flow and cutting performance, and it is even
harder to measure them. There seems to be no general agreement among authorities
on the subject, although this may be due in part to differences in basic definitions and
reference planes. It is beyond the scope of this text to argue the identification of the
functional angles, but a few examples will give the prospective tool designer a basic
understanding.
The functional angle of greatest importance in metal cutting is the true-rake angle.
It is generally defined as the slope of the tool face in the direction of chip flow with
respect to a radial reference plane passing through the point of the tool and the axis
of the rotating workpiece. The radial reference plane is established in space by the
axis of work rotation in single-point turning. Rotating cutters use an equivalent axis
of cutter rotation. The tool shown in Fig. 4.64 is used for the derivation of the true
rake-angle equation. Here the top view of a single-point tool with a sharp nose and a
side cutting-edge angle is shown. The basic reference plane passes through the tool
nose S and the axis of rotation. An additional reference plane is passed through any
point Q on the cutting edge (parallel to the basic reference plane), and triangle QPT
is drawn in the reference plane to correspond to the angles whose relationships are
to be determined. The equation of the true rake W will be determined as a function of
the side rake s, back rake b, and side cutting-edge angle e. Point P on the reference
plane drops to P2 because of back-rake angle b, and point T drops to T1, because of
the side-rake angle s. Thus PP2, the side opposite the true-rake angle W is
Design of Cutting Tools 279

Reference plane through Q


R R

s s

R
e T1
TT1 TQ
e
l b

PP1 P2 P P1 P2

R Q l R

l tan b
sin e l cot e

w R
R
P
l tanRw
sin e Q PT
P2 s R
R l cot e tan s
T1

Fig. 4.64 Derivation of the equation for true rake

PP2 = TT1 + P1 P2
From Fig. 4.64,
TT1 = l cot e tan s
P1 P2 = l tan b
l
PP2 = ____ tan W
sin e
Substituting,
l
____ tan W = l cot e tan s + l tan b
sin e
Solve for tan W:
tan W = cos e tan s + sin e tan b
A single-point cutting tool has a back rake of –5°, a side rake of
Example 4.20 8°, and a side cutting-edge angle of 30°. What is the true rake?

Solution: tan W = cos 30° tan 8° + sin 30° tan (–5°)


= 0.866 × 0.140 + 0.500 × (–0.087)
= 0.1212 – 0.0435
280 Tool Design

= 0.0777
W = 4°27¢
A second important functional angle is the angle of inclination, or the inclination of
the cutting edge with respect to the radial reference plane. The same component
angles that determine the true-rake angle (the side rake, the back rake, and the side
cutting-edge angle) also determine the angle of inclination with respect to the refer-
ence plane. The nature and escape of the chips depend upon the angle of inclination.
If the angle of inclination is positive, i.e., if point Q (Fig. 4.65) leads all other points
on the cutting edge when the tool is fed into the workpiece and the friction is low, the
chips flow easily outward. A negative angle will cause the chip to flow inward toward
the finished surface, marring it and (in all but the simplest tools) clogging the available
space.

Fig. 4.65 Angle of inclination

A formula for the determination of the angle of inclination developed by Max Kronenberg
is derived below
Let r = angle of inclination
Design of Cutting Tools 281

b = back-rake angle
s = side-rake angle
c = side cutting-edge angle
Then, from Fig. 4.64,
cos c = DQ/AQ and AQ = DQ/cos c
AB = AQ tan r = DQ tan r/cos c
CD = DQ tan b (from section YY )
From section XX
CG = BG tan s
but BG = AD = DQ tan c
CG = DQ sin c/cos c tan s
Also, from section XX
AB = CD – CG
DQ tan r/cos c = DQ tan b – DQ sin c/cos c tan s
Multiplying throughout by (cos c)/DQ gives
tan r = tan b cos c – sin c tan s
A lathe tool has a side rake of 25°, a side cutting-edge angle of
Example 4.21 30°, and a negative back rake of 10°. Since the finished surface
on the workpiece was poor, the tool designers were asked to
investigate.
Solution: tan r = tan a cos c – tan r sin c = – tan 25° sin 30°
= – 0.386 = – 21°6*
Since the inclination of the cutting edge has a large negative angle, the chip tends to
coil up in a tight spiral toward the finished surface, scraping it and generating exces-
sive friction.
From a practical viewpoint, the nomograph in Fig. 4.66 is of more use to the tool
designer in the field. It enables the designer to determine true rake and inclination by
connecting variables with a straightedge or by drawing a straight line.
To determine the true rake of our previous example (back rake –5°, side rake of 8°,
SCEA of 30°), draw line R from the –5° point on the back-rake scale to the +8° point
on the side-rake scale. At its intersection with the 30° vertical line from the black
SCEA scale read approximately +4.5°.
To determine the angle of inclination in the previous example, draw line C from the
+25° point on the side-rake to the –10° on the back-rake scale. At its intersection with
the 30° vertical line from the gray SCEA scale read approximately –21.6°.
The nomograph in Fig. 4.66 readily shows the effect of the side cutting-edge angle on
the magnitude of true rake and inclination when the back rake and side rake are held
constant. They can easily be changed by grinding a different side cutting-edge
282 Tool Design

angle on the tool. It further shows that true rake is always positive when both side
rake and back rake are positive, and negative when both are negative. This is to be
expected, but many machine-tool people do not realise that true rake can be posi-
tive when either side rake or back rake is negative. This enables the tool designer
to specify a tool with a negative rake angle and still have the effect and efficiency of
positive rake angles.

Fig. 4.66 Chart for determining true rake and inclination (Cincinnati Milacron, Inc.)

Types of single-point tools The types of tool commonly used for single-point cutting
operations are the solid, the brazed-tip, the long indexable insert, and the throwaway
indexable insert types. Many variable factors determine the proper selection of tool
type, of which the workpiece material, the machine tool itself, the type of set-up, the
amount of material to be removed, the power available at the machine, and the kind
of operation are a few.
The solid-type tool bit is one made entirely of the same material, as opposed to the
insert type, in which an insert is brazed or held mechanically to the shank of another
material. High-speed and cast-alloy tools are commonly of the solid type. Solid tools
may be purchased already ground, but as a general rule they are ground to the
required geometry by the machinist or machine operator.
Cost is the prime factor for determining the use of solid-bit tools. The materials used
in solid-bit tools are less expensive. They are also used for special-ground form tools
since the material is easier to grind to special shapes. High-speed steel and cast-alloy
tools may be ground with conventional grinding wheels, but carbide requires the use
of a diamond grinding wheel.
Design of Cutting Tools 283

Solid-bit tools can be purchased as squares, rounds, rectangles, and special shapes.
They may be mounted on the machine tool directly in the tool post or block, but the
smaller sizes usually are held in a toolholder, as shown in Fig. 4.67. Note that the
solid-bit tool is tipped up to an angle of 15 to 20°. This provides back rake without
grinding it into the tool bit. This type of toolholder requires that additional front relief be
ground in the amount the tool bit is tipped upward. For example, in order to provide a
front relief angle of 8° in a toolholder that held the tool bit at an angle of 16°, it would
be necessary to grind 24° front relief on the bit itself. The end result (when the tool
bit is mounted on the machine in the toolholer) would be 8° front relief and 16° back
rake.

(a) (b)

(c)

Fig. 4.67 Toolholders for solid bit tools (a) Left-hand offset, (b) Right-hand offset and
(c) Sstraight-shank turning toolholder (Armstrong Bros. Tool Co.)

Brazed-tip tools have the cutting insert held in the tool shank by silver brazing (solder-
ing). Tungsten carbide is almost exclusively used as the tool material for brazed-tip
tools. There are three basic methods generally used for brazing the carbide tips to the
tool shanks: torch, gas burner, and induction heating. In all three methods, a pocket
is machined in the shank material to fit the shape if the carbide insert. Brazing with
a torch consists of using an oxyacetylene torch as a heat source to melt the braz-
ing alloy. This method is used in smaller shops where brazing equipment is limited.
High-frequency induction heating is used as the heat source by larger industries and
manufacturers who produce brazed-tipped tools in quantity. An advantage of induc-
tion heating is that the heat is generated within the tool shank and the carbide tip. It is
much faster than torch or gas-burner brazing.
Figure 4.68 shows a typical brazed tool. Many sizes, styles, and grades are available
as standard items, or the carbide inserts are available for brazing in the customer’s
own plant. Brazed tools have distinct advantages, the first being that they are low in
unit first cost. They do not require the space that a mechanical toolholder does, and
therefore can be ganged or used in tooling devices where crowded conditions exist.

(a) (b)

Fig. 4.68 Typical standard brazed tools (Kennametal, Inc.)


284 Tool Design

They are also used for the construction of special tools with formed cutting edges that
are not available as standard items.
Brazed tools have their disadvantages. Carbide has a lower coefficient of expansion
and when heated expands approximately one-half as much as steel. This means
that in a brazed tool assembly, the steel will lengthen considerably more than the
carbide. Upon cooling after brazing, the steel will contract more than carbide, causing
thecarbide to be under stress. Premature tool failure may be the end result if the tool
is poorly designed or carelessly brazed. Brazed tools also require considerable time
to change in the machine. In addition, they have only one cutting edge, which must
be reground before it can be reused after wear.
Long indexable insert tools, sometimes called
on-end or slug-type tools, consist of a mechan-
ical toolholder and an indexable insert with
tops and bottoms that can be reground to the
correct geometry (see Fig. 4.69). The insert is
approximately 38 mm long and is held verti-
cally with the top of the insert tipped forward
and in the direction of feed a few degrees to
Fig. 4.69 A long indexable-type tool
provide the necessary relief. Regrinding is
(VR/Wesson Company)
done by grinding the ends of the insert normal
to the sides only enough to clean up the cutting edge. The toolholder has provisions
for readjusting the insert after resharpening. Inserts are available in square, rectangu-
lar, diamond, round, and triangular shapes. When chip breakers are necessary, they
must be ground into the tool.
The throwaway or disposable insert is one
of the latest developments in single-point
tools (see Fig. 4.70). The term throwaway
or disposable refers to the cutting-tool insert
which is mechanically held in the toolholder.
The inserts are purchased ready for use
and, depending upon the shape and insert Fig. 4.70 A heavy-duty disposable in-
geometry, have a number of cutting edges sert-type cutting tool (Wendt-
that can be indexed into position. When all Sonis, Division of United
cutting edges are used, the insert is dis- Greenfield Corporation)
carded and not resharpened. This approach
to tooling completely eliminates tool grinding and resharpening. Inserts are available
in triangular, square, round, diamond, parallelogram, and button shapes. All edges of
the square and triangular inserts can be used, and with negative rake inserts, both
top and bottom edges are to be used. Thus, with a negative triangular insert, there
are six cutting edges per insert. Positive rake tools are indexable on one side only
(see Fig. 4.71).
In addition to the elimination of tool grinding with throwaway insert-type tools, brazing
and its problems are eliminated. Since the inserts can be indexed quite accurately,
the time required to change tools is greatly reduced. Since grinding and tool-changing
time are reduced, the machine operator need not baby the cutting tool to make it last
longer. Feeds and speeds can be increased to further increase production. When a
cutting edge becomes worn, it takes only a few seconds to index a new cutting edge.
The machine tool is made to work to its maximum capacity.
Design of Cutting Tools 285

Negative insert; both


top and bottom edges
may be used
– –

Insert seat –

Position insert; only +


top edge may be used

0° +

Fig. 4.71 Relationship between positive and negative disposable inserts (Clamping mecha-
nism is not shown)

The toolholders for throwaway insert-type tools are designed for extreme rigidity.
Basically, all toolholders have a solid base with a pocket or nest to receive the inert
and hold it into position. The material used in the toolholder shank is some type of
tool steel which is heat-treated from RC40 to 45 to resist clamping screw damage and
chip erosion.
The nest or pocket is one of the most critical areas of the toolholder in that it must be
flat to provide proper support and avoid inducing strains in the insert. It must be accu-
rate to within 1° at the corner angles and must have straight sides. Limits of 0.013
mm are held in the geometry of the pocket to ensure accurate indexing in addition to
keeping insert movement to a minimum. The inside corner of the pocket is generally
relieved to provide chip clearance.
A carbide seat or pad is usually provided as a support for the insert, although some
companies do not provide seats in their smaller toolholders. The purpose of the seat
is to transfer heat away from the cutting edge and to transfer cutting forces over a
wider area of the pocket. It also partially protects the pocket in the event of insert wipe
out or chipping. Like the pocket, seat flatness is a critical requirement. It is said that
an out-of-flatness condition of as little as 0.05 mm can result in insert breakage. The
sea is usually attached in the nest for convenience and to prevent its shifting or loss
during insert indexing or removal. The seat material is usually an impact resisting
grade of carbide because of its high compressive strength, although high-speed steel
and other materials are also used.
Various designs of mechanical insert clamping devices have been used. The main
function of the clamping mechanism is to hold the insert securely in position, although
in some designs it also serves as a chip breaker. The clamping mechanism should
carry very little of the load caused by cutting forces. The insert in most styles of tool-
holder is held in the pocket by the cutting forces, and the load on the clamp is quite
286 Tool Design

light. This has been demonstrated by actually removing the clamping mechanism
after the cutting edge is in the cut.
The shifting of the insert within the pocket is sometimes encountered with tracing
operations. Metal-cutting forces are constantly changing direction and magnitude dur-
ing tracing operations, and there is sometimes a tendency to twist or pull the insert
out of the pocket. This tendency is aggravated when the tracing operation requires the
cutting tool to cut on the back side of the tool. Difficulties of this nature can generally
be overcome by using toolholders especially designed for tracing operations (see Fig.
4.72).

Clamp screw
Clamp screw
Plain bridge clamp
Cone-point clamp

Chip breaker

Plain insert N/P inserts

Lockpin Shim pin


Shim Shim

Lock screw

(a) (b)

Fig. 4.72 Toolholders designed especially for tracing and profiling (a) Exploded view showing
holder components. The chip-breaker plate is not required when using inserts with
built-in chip control. For tool set-ups where side locking screw is inaccessible, an
alternate clamping method is shown in (b). (b) Alternate insert-clamping method
for 35 and 55° holders. The cone-point clamp and shim pin must be ordered
separately. (Kennametal, Inc.)

The mechanical clamping mechanisms employed by


toolholder manufacturers vary in design, but one of
three different clamping methods is generally used.
They are the screw, the bridge, and the pin type. The
screw type shown in Fig. 4.73 holds the insert in place
in the holder with a screw. It is used for jobs where the
design or size of holder does not permit use of other
types of clamping mechanism. The insert is released
from the top of the tool, which is an advantage when Fig. 4.73 Screw-type clamping
the tool is held in such a way that the bottom of the mechanism (Ken-
tool cannot be reached. The insert locates against the nametal, Inc.)
recess shoulders, as the holder screw hole is slightly oversize, allowing the insert
to “float”. Typical applications for these tools are small boring bars and automatic
tooling. A disadvantage of the screw-type holder is that the screw sometimes sticks
because of heat and minute chips that work their way into the threads. The wrench
socket will also fill with fine metal particles, making it difficult for wrench insertion.
Design of Cutting Tools 287

The bridge-type clamping mechanism is easily recognised by a bridge-type clamp


that directs clamping forces downward by means of a screw located in the centre
of the clamp (see Fig. 4.74). The bridge varies in size and shape, depending upon
the manufacturer, but the principle is similar in all designs. A chip breaker is gener-
ally used with the bridge clamp and may either be a permanent part of the bridge or
a replaceable separate chip breaker that is similar to the cutting insert but smaller.
Some manufacturers secure this type of chip breaker to the bridge to prevent loss
during insert indexing.

Fig. 4.74 Carbide toolholders utilising bridge-type clamping mechanisms (Wendt-Sonis,


Division of United Greenfield Corporation)

The main disadvantage of bridge-type clamps is that the clamp and the chip-breaker
mechanism result in a bulky or club head, which sometimes makes it difficult to use
in close quarters. On large machines where extremely heavy cuts are required, the
heavy chip may wipe off the chip-breaker and clamping mechanism.
The pin-type clamping mechanism is similar in design to the screw type in that the
insert features a hole in the centre. A pin is used to wedge the insert against the
shoulders of the pocket. This design does away with clamps, screws, plates, and
chip-control devices from the top of the tool, thus eliminating many of the disadvan-
tages of bridge clamps. Shorter tool overhang allows heavier cuts with smaller shanks
and minimises the possibility of deflection and chatter. Extremely neavy chips slide
over the top of the insert without obstruction. Chip control, when needed, is accom-
plished by means of a groove-type chip breaker along the edge of the insert. Although
the chip breaker has been designed to give a wide range of chip control within depths
of cut and feed rates for which the inserts are normally used, it cannot be changed
to meet specific cutting conditions. Figures 4.75 and 4.76 shows toolholders utilising
pin clamps.
An unusual type of clamp for positive-rake toolholders is shown in Fig. 4.77. In addi-
tion to wedging the insert against the shoulders of the pocket, the pin pulls the insert
down against the seat. This design has proved quite popular, especially with smaller
machines or workpieces that lack the rigidity for successful negative rake turning. It is
ideal for machining thin-walled parts, thin webs, and long slender shapes and tubes.
288 Tool Design

Carbide
Shim Lock screw
insert

Shim lock
Lockpin

Fig. 4.75 Simplok pin-type clamping mechanism (Carmet Company)

Chip control with single-point tools The


removal of metal in the form of chips cre-
ates the problem of transferring the newly
formed chip away from the cutting area.
Ideally, this is accomplished by causing
the chip to break into small segments in
order to allow it to fall away from the cut-
ting tool into the chip pan below.
Removing chips produced from a mate-
rial that forms a segmental chip when
machined poses no problem. Chips
formed at speeds in the high-speed-steel
cutting range (23 to 61 m/min) usually do Fig. 4.76 Carb-O-Lock eccentric pin-type
not pose a problem either, as the chip has clamping mechanism (Metal-
a natural tendency to curl, break, and fall lurgical Product Department,
into the chip pan. A favourable chip can General Electric Company)
be produced at these cutting speeds by
simply increasing the feed in relation to the depth of cut or by slightly decreasing the
rake angle of the tool. Many machinists choose a certain depth of cut then increasing
the feed until an acceptable type of chip is produced. If the feed cannot be increased
Design of Cutting Tools 289

Fig. 4.77 Positiverake Carb-O-Lock pin-type clamping mechanism (Metallurgical Product


Department, General Electric Company)

sufficiently because of limitations imposed by the machine or work set-up, the pro-
cess is repeated with a smaller depth of cut.
Newer cutting-tool materials have brought about a drastic increase in cutting speeds.
The chip material becomes more plastic with the elevated temperatures produced
at higher cutting speeds and has a tendency to flow rather than break. A continuous
chip is formed which may wrap around the tool post, the workpiece, the chuck, and
maybe even the operator. In addition to endangering the operator, such a condition
causes valuable production time to be lost for stopping the machine and removing
the entangled mass of continuous chips. Such a mass is also bulky and causes chip
disposal and storage problems. It may also mar the newly finished workpiece surface
and interfere with the flow of coolant. Automatic machine tools may not function prop-
erly because of chip interference. Therefore, at high cutting speeds, it is imperative
that the chip be controlled and broken in some manner.
The basic method of breaking the chip is to place an obstruction in its path. This
is generally done by grinding a step or groove on the face of the tool or by using a
mechanical chip breaker. This obstruction interferes with the natural flow of the chip
and causes it to take on considerable curvature. The obstruction is so placed that it
forces the curved chip to collide with either the uncut portion of the workpiece or the
flank of the tool below the cutting edge and produces chip fracture at short and regular
intervals.
There is an apparent paradox, in that efforts are always being made toward improving
the flow of the chip across the face of the tool rather than obstructing it. An obstacle
deliberately placed in the path of the chip would obviously impose additional defor-
mation and absorb extra energy. The trick is to design the obstacle to restrict the
natural flow of the chip as little as possible. Sometimes a slight obstacle is all that is
necessary to cause the chip to curl and break. Too large an obstacle may cause a
tight “crowded” chip that imposes heavy forces on the cutting edge and the obstacle
itself. Mechanical chip breakers features an adjustment which allows the operator to
vary the distance between the obstacle and the cutting edge and make it possible to
deform the chip just enough to cause it to break and yet only slightly retard the natural
flow of the chip. This type of chip breaker can be used successfully on all types of
materials, while ground-in chip breakers may have to be altered before machining a
different material.
290 Tool Design

Chip breakers are generally classified as ground-in, mechanical, and pressed-in


types and are generally used in brazed tools and regrindable inserts. They consist
of shallow step or groove ground along the cutting edge to deflect the chip back into
the workpiece or against the side of the tool, where it is broken. A diamond wheel is
required to grind the step or groove in carbide cutting-tool materials in order to hold
the dimensions. A ground chip breaker is not adjustable, and therefore it must be
used on the particular job it was designed for.
The basic types of ground-in chip breakers are referred to as the angular, parallel, and
groove. The angular type shown in Fig. 4.78, is often used for general turning since it is
the easiest to grind. The angle of the step relative to the cutting edge is usually from 8
to 10°. Since the groove is narrower at the back, the chip is curled tighter at the outside


0.4
0.8
0.25 R.
0.5

Fig. 4.78 Angular-type chip breaker for general use in continuous cuts on concentric work-
pieces (Wendt-Sonis, Division of United Greenfield Corporation)

diameter than near the point of the


tool. This tends to curl the chip away 1.6
from the workpiece. However, in deep 45°
cuts, the back of the chip tends to be
“too tight” and may cause excessive 0.4 0.8 R.
heat and pressure. This same factor 0.8 0.25 R.
may be used advantageously in turn- 0.5
ing operations when the depth of cut
varies, provided, it is not too deep. As
the depth of cut increases for a given
feed in some materials, the chip will
become stronger and will curl rather
than break. The angular breaker with Fig. 4.79 A 45° chip breaker for light or fin-
its diminishing width will crowd the ishing cuts having maximum depth
chip tighter with increasing depth and of 0.8 mm (Wendt-Sonis, Division
cause it to break. of United Greenfield Corporation)
The chips produced by light finishing cuts of 0.8 mm or less are sometimes effectively
controlled by an angular chip breaker ground at 45° (see Fig. 4.79). Tools that have a
Design of Cutting Tools 291

large nose radius require the angular


chip breaker to be modified as shown
in Fig. 4.80 because the chip-breaker
width must be wider than the nose
radius at the point of the tool. The
second angle is used to prevent the
top of the breaker shoulder from hit-
ting the workpiece.
The ground-in parallel-type chip
breaker shown in Fig. 4.81, is used
for cuts extending to a 90° shoulder.
It directs the chips against the work,
Fig. 4.80 Chip breaker for tool having a large
causing it to break, and is satisfactory
nose radius and used with heavy feeds
for deeper cuts.
(Wendt-Sonis, Division of United
The width of ground-in step chip Greenfield Corporation)
breakers depends to a great extent
upon the feed rate. Table 4.17 give suggested widths of various feed rates.
The ground-in groove breaker shown in Fig. 4.82, consists of a narrow groove ground
so that a narrow land is left between the groove and the cutting edge. This type of
breaker usually produces a curl rather than breaking the chip into small pieces. This
kind of chip formation is often preferred for automatic lathes and screw machines,
where short chips would interfere with the mechanism. A general rule for dimensions
is to provide a land width of 1 to 1.5 times the feed, a groove width of three to four
times the feed, and a groove depth of 0.25 mm.
The mechanical or clamped-on chip breaker utilises a small shim of carbide with a
bevelled edge. The shim is clamped firmly to the tool face with the bevelled edge
toward the cutting edge. Two styles are shown in Fig. 4.83. One style features a com-
bination clamp and chip breaker, where the carbide shim is brazed into the clamping
mechanism. The advantage in this case is the one-piece construction. There is no
adjustment with this style, but many companies have the combination chip breaker
and clamp available in different widths in order to meet the demands of cutting con-
ditions. In other words, if the chip is too tight, the chip breaker and clamp can be
removed and replaced with one that has the carbide shim further away from the cut-
ting edge.

Table 4.17 Recommended widths of chip breakers for various feed rates and depths of cut
Feed, mm/rev
Depth of cut, mm 0.2-0.32 0.33-0.49 0.45-0.56 0.57-0.70 0.71-0.81
Width of chip breaker, W (mm)

0.40-6.19 1.59 1.2 2.38 2.78 3.18


1.59-6.36 2.38 3.18 3.97 4.37 4.77
7.94-12.71 3.18 3.92 4.77 5.16 5.56
14.30-19.07 3.97 4.77 5.56 5.96 6.36
NOTE: A chip-breaker depth of 0.5 mm is satisfactory for most types of steel.
SOURCE: Wendt-Sonis, Divisions of United Greenfield Corporation.
292 Tool Design

1 TO 1 1 ¥ Feed
2
W

0.4
0.8 0.25 R.
0.5

Fig. 4.81 Ground-in parallel-type chip breaker Fig. 4.82 Ground-in groove-type
for cuts extending to a 90° shoulder chip breaker
or for use on eccentric workpieces
(Wendt-Sonis, Division of United
Greenfield Corporation)

The second style utilises a separate carbide shim held in place by the clamping mech-
anism. Various widths of shims are available to meet the various cutting conditions. A
disadvantage of this style is that the extra loose parts can be lost in the chip pan.
A modification of the second style is shown in Fig. 4.84. In this design the carbide
shim is adjustable by means of a screw thread. This permits close control of the chip
without dismantling the toolholder. Other designs of adjustable chip breakers utilise
serrated clamps and shanks to vary the chip-breaker width.
The pressed-in, or moulded chip breaker is shown in Figs 4.85 and 4.86. It has the
appearance of the ground-in groove, but instead of being ground, the groove is
moulded in place at the time of insert manufacture. Inserts utilising the pressed-in
groove are generally designed for pin-type toolholders. In order to give maximum

Fig. 4.83 Basic types of mechanical or clamped-on chip breakers


Design of Cutting Tools 293

cutting strength in the cutting edge, a nar-


row negative land is provided between the
cutting edge and the groove. The shape of
the groove can be varied in order to obtain a
positive rake angle in a negative rake holder
and still used negative rake inserts. Studying
the groove shapes in Figs 4.85 and 4.86,
shows that they could not be economically Fig. 4.84 Toolholder featuring an ad-
ground into the insert and must be put in dur- justable chip breaker (Metal-
ing its manufacture. lurgical Product Department,
A remarkable feature of the pressed-in chip General Electric Company)
breaker is that it will produce acceptable
chips over a wide range of materials, feeds, speeds, and depths of cut. The groove
deflects and curls the chip into acceptable form during light feeds, while on heavier
feeds, the chip skips the groove and acts like a regular negative rake insert. In other
words, when the chip becomes so heavy that it is not affected by the groove, it will
break anyway. The positive rake obtained as a result of the groove ground behind the
narrow negative land may range from 5 to 20°, depending upon the manufacturer. It
gives good chip flow on some of the soft and gummy materials, while the negative
land gives added strength to the cutting edge typical of negative rake inserts.

Fig. 4.85 Toolholders utilising the pressed-in, or moulded, chip breaker (Metallurgical
Product Department, General Electric Company)
294 Tool Design

Boring tools In the general sense, boring


tools may be classed as single-point tools;
however, this is not true in all cases since
some types of boring tools have more than
one cutting edge. Perhaps a better descrip-
tion would be to say that boring tools are
always associated with the process of
internal machining because a boring oper-
ation consists of enlarging to accurate
dimensions a previously drilled, cored, or
punched hole. This fact alone raises prob-
lems generally not associated with external Fig. 4.86 Carbide insert utilising 15°
turning because of the concave nature of positive-rake moulded chip
the surface being machined. grooves in negative-rake tool-
Figure 4.87a shows the relationship of an holders (Kennametal, Inc.)
external turning tool to a convex workpiece
surface, while Fig. 4.87b shows the relationship of an internal boring tool to a con-
cave surface. Both tools have a zero back-rake angle with respect to a vertical plane;
however, if a line is drawn from the point of the tool tangent to the workpiece circle,
it is easily seen that the convex shape in the turning operation gives a positive rake
effect, while the concave shape in the boring operation gives a negative rake effect.
The clearance angle is affected in the same manner. Greater clearance is provided
with an external turning operation, while less clearance is available with internal bor-
ing operations. These problems are compounded on smaller bores.
The problem of chip disposal is more important in boring operations because there
is no natural exit through which they can pass. Long bores of small diameter pose
a particular problem. Chip removal is usually accomplished by using an air blast, by
liquid coolant flow, or by boring in a vertical position so that chips can fall through the
bore.

Negative rake
effect

Positive rake
effect

Tool Tool

Greater relief Less relief


Relief becomes
smaller as bore
diameter decreases
(a) (b)

Fig. 4.87 External turning compared with internal boring


Design of Cutting Tools 295

Boring operations generally require a longer tool with a higher length-to-diameter ratio
in order to reach into the bore. The stability of the tool may be reduced to the point
where chatter may occur. Such a condition may call for additional supports or special
techniques to attain suitable efficiency in the boring operation (see the subsection on
chatter and vibrations in Section 4.5).
Types of boring tools The selection of boring bars depends to a large extent on
the type of work being performed. A shop or company doing a variety of small lot
jobs would need to select bars that could be used on a large number of different
workpieces. In this case, the ease and rapidity of adjustment, the interchangeability
of different shaped cutters, and general versatility are important. On the other hand,
long production runs call for a heavy bar suitable for maximum feeds and speeds,
which may incorporate several tools precisely located to perform several machining
operations at one time. The production bar may be especially designed to machine
only one workpiece. The following discussion will describe the various types of bars
and boring tools used in industry today.
Stub boring bars are those which are supported only at one end and are used to bore
relatively short holes. In this case, the bar acts as a cantilever beam when in action,
since it receives a force at one end and is supported at the other. Stub boring bars
should be as short and rigid as possible for this reason.
Stub bars used in small diameter holes are generally forged or preformed from the
solid. Figure 4.88 shows a forged example. In recent years, this type has been los-
ing favour and is being replaced by commercial performed bars (see Fig. 4.89), a
type designed to produce a constant clearance angle on both the front and side cut-
ting edges. Resharpening is accomplished by simply grinding the top face. Repeated
reshapings will not alter the clearance or change the form of the cross section. Bars of
this type are made to bore holes as small as 1.6 mm in diameter. Figure 4.90 shows
the basic tool geometry of this type of bar. The importance of correct setting is shown
in Fig. 4.91, as this type of bar is designed to work on. The set-up must be as accurate
as possible. Raising or lowering the cutting edge by rotating the tool decreases or

Top

Workpiece

Back rake 5°
Side rake 20°

Front
Side relief 10°

Clearance 20°

Fig. 4.88 Forged boring cutter


296 Tool Design

Fig. 4.89 Preformed boring tools designed to produce a constant clearance angle on both
front and side cutting edges (Bokum Tool Company, Inc.)

increases the clearance accordingly. When


a change in back rake is required, it should
be ground into the tool and not achieved by
rotation.
Stub bars for larger holes may utilise a tool
bit held the bar by a square slot and set-
screw. Bars of this type are generally used
for general work on the engine lathe and are 30°
intended for fairly light work under the direct
control of a skilled machinist. The square slot
may be at 90° to the bar axis on one end for 12° rake
thought boring, while the other end may con-
tain a 45° slot for boring blind holes.
Stub boring bars are also commercially avail-
able in the throwaway insert type. Figure 4.92
should two different styles. The bar in a fea- Fig. 4.90 Basic tool geometry of
tures a micrometer adjustment built into the preformed boring tool
head of the bar, whereas b is a standard for
straight boring.
Line boring bars, sometimes referred to as pilot boring bars, are those which extend
through the workpiece and are supported in more than one place. Usually, this support

(a) (b)

Fig. 4.91 The importance of correct setting of preformed boring tools. (a) Correct: tool is
designed to work on centreline (b) Incorrect: lowering tool centreline destroys
relief, causing interference
Design of Cutting Tools 297

(a) (b)

Fig. 4.92 Stub boring bars of the throwaway insert type (VR/Wesson Company)

is provided at both ends of the bar. A general purpose line boring bar is character-
ised by a large length-to-diameter ratio and may have provisions for standard tool
bits, block tools, boring heads, or tool cartridges. It may contain one tool for a single
machining operation or utilise several tools to perform multiple boring, facing, counter
boring, and chamfering operations. Bars with stepped diameters are sometimes nec-
essary to accomplish multiple machining operations.
Figure 4.93 shows one method of adapting a line boring bar to a standard engine
lathe when through boring circular work that can be conveniently held in a standard
chuck. Workpieces not conveniently held in the chuck are sometimes clamped directly
to the cross slide of the lathe. This practise is limited to larger lathes, where the slots
or other provisions for clamping the workpiece are incorporated into the cross slide.
In this case, the boring bar is held and driven by the lathe spindle, and outboard bar
support is provided by the lathe tailstock centre.

Jaw
Chuck Workpiece
Tool holder
Lathe spinale

Boring bar
Pilot bushing Cutting tool

Fig. 4.93 Method of adapting line boring bar to a standard lathe

Although line boring bars may be used on the lathe other standard machine tools,
they are most commonly associated with the horizontal boring machine. This machine
has provisions for clamping for moving the workpieces in order to bore more than one
hole in any one setting or clamping of the work piece. An adjustable outboard support
298 Tool Design

is used to support the bar after it has been positioned the bore of the work piece.
Figure 4.94 shows various designs of line boring bar. Bars a and b are simple and
represent the type used for small-lot production. Bar c gives a good idea of the pos-
sibilities of special multiple-cutter bars designed for high-production work. Selecting,
designing, and supervising this type of machining setup forms a large part of the tool
designer’s or tool engineer’s work in connection with boring bar. Tool room and turret-
lathe boring bars are generally purchased complete, and the machinist needs very
little help from the plant tool designers in this connection; however, special multiple-
cutting bars need careful study based on experience. The design is generally worked
out through the co-operation of the tool designers of both the purchasing and selling
companies. The drawings need little explanation, but it should be noted that besides
setting cutters to machine several diameters simultaneously, it is possible to rough-
bore a hole and then follow this operation immediately with a finishing cut. The time
saved by this procedure is apparent. In Figure 4-94c, note also the hardened insert as
a protection against chip wear and the ducts to carry coolant to the cutters.
The accuracy of line boring bars depends, to a large extent, on the fit between the
pilot and its bushing. If the clearance is too small, the pilot may seize in the bush-
Graduated dial Flat
Hardened and ground pilot

Cam-type lock
screw
(a)

Pilot Expanding block- Shank


type double cutters

(b)

Hardened and Rough bore Finish bore Thread bore Counterbore


ground pilot

Stellite chip Lubricant line Face and


protection chamfer
(c)

Fig. 4.94 Line boring bars


Design of Cutting Tools 299

ings. If the clearance is too large, it may be impossible for the operator to hold the
workpiece to the tolerance specified.
The actual determination of clearance is a matter of judgment based on experience.
The fit between a boring bar and its bushing can be much closer than that allowed
in the general case of running fits. In the first place, the action of the boring bar is
intermittent. This reduces the possibility of overheating with the degree of seizure.
Furthermore, boring-bar bushings are usually flooded with a coolant, which, although
having only moderate lubricant possibilities, does help guard against failure caused
by metal-to-metal contact or excessive thermal expansion. Many boring machine
have an outboard bushing whose outside diameter is tapered and can be adjusted
to the proper clearance at the beginning of the machining operation. The machine
operator adjusts the bushing so that the bushing housing is warm to the touch during
continuous operation.
In ultra precision work and high-speed boring, it is possible to use rotary bushings run-
ning in preloaded ball bearings. Deflection is reduced almost to zero by the principle
of preloading, a principle extensively applied in the machine-tool industry. It is accom-
plished as follows: the diameter of the bushing or outer race, making the equivalent of
a press fit. Since most of the deflection of a beam occurs during the initial loading, and
since the structure of ball bearings is similar, the actual deflection shows a marked
decrease with an increase in load. Thus, if the large deflection due to the first loading
is eliminated by preloading before the working load is applied, the working deflection
is considerable reduced and of course the carrying capacity of modern ball bearings
is such that they can easily carry the two loads.
Incorporating rotary bushings in the design of a boring bar not only promotes rigid-
ity but reduces wear. This is especially advantageous for long production runs at
high speed on large horizontal boring machines. The abrasive action of metallic dust,
especially cast-iron dust, acts, as a very efficient lapping compound when mixed
with cutting fluid. Ball bearings are by no means so vulnerable to this action as plain
bushings.
For high speeds exceeding 29 m/min on the bearing diameter, it is recommended that
ball-bearing bushings be used, especially when using carbide cutting tools.
Hardened steel or bronze insert strips placed axially in the pilot sections of bars
are sometimes used when rotary bushing cannot be. The bar is first spline-milled
or hobbed, and the strips are closely fitted into the slots to form the pilot surface.
Countersunk screws are used to hold the strips in place, and as wear occurs, the
strips are shimmed up and reground to proper size. Cleance between the strips and
the bushings can be as close as 2.5 mm for bars up to 75 mm in diameter and 50
mm for those over 75 mm. For best results, the wear strips should constitute about
one-third of the circumference in bearing surface and should project approximately
1.6 mm above the bar diameter.
Multiple-cutter boring heads are used for larger bores where machine rigidity and
access to the work surface permit. A boring head reduces cutter overhang and allows
greater feed. It may be mounted directly to the spindle by an integral shank or stub
arbor, or it may be mounted directly on a line bar with a key between the bar and the
hub of the boring head.
The radial-cutter boring head shown in Fig. 4.95 acts somewhat like a reamer when
cutting. The cutters are placed radically around the bar and are adjusted by sliding
300 Tool Design

them endwise in tapered slots and locking them by clamping plates as shown. This
type of head is suitable for boring short holes where fast action in essential and where
it is possible to dispose off the chips without great difficulty.

Inserted cutter Wedge block Chip clearance fluting

Fig. 4.95 Radial-cutter boring bar

Figure 4.96 shows a simple boring head


Micrometer screw-on-type
designed to prevent the excessive cutter boring cartidge Head keyed to bar
overhang that would be necessary with a
conventional mounting. The head may be
located anywhere along the bar because it
is demountable. Multiple operations can be
performed by mounting several heads on
one bar. This type of head is well suited for
mounting block- or cartridge-type cutters
because of the increased area provided
for mounting. Quite often a boring head of
this type is designed with two or three cut-
ters set for roughing cuts and one single
cutter set for a final cut. In this case, each Fig. 4.96 Demountable boring head
cutter would have an individual micrometer
adjustment for each of the settings in relation to the others.

Boring-bar cutters and adjustments A large number of designs for holding and
adjusting cutters for boring operations have been developed over the years. The
object of all cutters of this type is to secure rigidity and accuracy of cutter setting in the
simplest manner possible. A simple type used extensively on turret lathes is shown
in Fig. 4.97. It will be readily seen that the cutter can be adjusted to bore different
diameters by means of three screws while rocking the cutter up and down to obtain
the size desired. A fourth screw on the side provides a means for final clamping of the
cutter once it is adjusted.
A similar type is used on jig-boring machines, vertical milling machines, and radial drill
presses. It is often referred to as a boring head or wobble head. The adjustment for
the bore diameter is made by turning the lead screw (see Fig. 4.98), which moves the
slide in the V block and changes the eccentricity of the stub boring bar.
Figure 4.99 shows common methods of holding and adjusting standard rectangular
tool blanks in stub and line boring bars. A single screws is used if adjustment of the
tool within the bar is not necessary, and two screw are used when such an adjustment
is desirable. One screw is to adjust and back up the cutter, and the second is to lock
the cutter into position. The locking screw can be placed on the side of the bar or at
Design of Cutting Tools 301

Fig. 4.97 Adjustable boring toolholder used on turret lathes (Warner & Swasey Co.)

Gib screw
Shank
Gib

Cutting
edge Cutting
setscrew Graduated dial

V-block Lead screw


Enlarged view
of cutter profile

Screw Washer

Fig. 4.98 An adjustable boring, or wobble, head


302 Tool Design

the end, whichever is most convenient. One locking screw is generally used on small
bars because of space limitations. The tool designer must consider the convenience
of operation, clearance room and interference with chips or parts of the workpiece
when deciding where to locate these screws.

Adjustment
(a) (b) (c)

(d) (e)

Fig. 4.99 Methods of holing and adjusting standard rectangular tool blanks in stub and line
boring bars

The square hole containing the cutter can Adjustment


be broached through the bar and counter
bored and threaded for a large adjusting
screw. However, if the cuts are heavy, it
may be better to broach the square hole
to a blind stop and drill a round hole much
smaller in diameter than the cutter cross
section. This avoids weakening the bar in
the outside fibres that resist the bending
Locking screw
action.
Often it is more convenient to drill and ream
a round hole in the boring bar and press a
sleeve containing a square hole as shown
in Fig. 4.100. Sleeves containing square
holes are available as a commercial item
from tooling specialty companies. Standard broached sleeve pressed
into a blind hole
Figure 4.101 shows s simple method of
fastening a cutter to a boring bar that can Fig. 4.100 Use of a standard broached
be used where a variation in hole diameter sleeve to hold a boring cutter
is unnecessary. The cutter is located by in a line boring bar
Design of Cutting Tools 303

means of shoulders and is held firmly in place by Rectangular


Wedge slot
a wedge. This cutter has the advantage of having
two cutting edges. Its disadvantage of having no
adjustment is largely offset by the fact that this
type of cutter blade can be inexpensively made
from flat stock. Shoulder

Double cutters of this type are most often used


for roughing cuts, since the number of regrinds
for a given bore is limited, especially when the 2° wedge
tolerances are close. To ensure that both cut- angle Cutter
ting edges have the same depth of cut, the cut-
ter is first locked in the bar and the unit mounted Fig. 4.101 Wedge-type holder
between centres of a tool and cutter grinder. After the first edge is ground, the bar is
turned 180° and the second edge ground at an identical radius from the bar axis.
The illustration shows the taper cut in the bar. While this method functions satisfacto-
rily, machining costs will be reduced if the slot is broached out square and the taper
placed on the cutter.
A two-bladed boring bar used extensively for machining cast iron is shown in Fig.
4.102. Figure 4.103 shows a convenient bar for light work, such as for cutting brass
or aluminium. The cutters are adjusted by sliding them lengthwise and locking them
in position by driving home the taper pins.

Adjustment screw Lengthwise adustment

Clamping screw Taper pin

Fig. 4.102 An adjustable boring toohol- Fig. 4.103 Boring toolholder for light
der suitable for cast iron work

Boring-tool designs of the preceding types have been used successfully in the past
but under present manufacturing conditions they have one disadvantage: they
require time and skilled toolmakers to manufacture. When commercial boring tools
were not available, there was no other solution, but now commercial tools are eas-
ily purchased. Standard bars and cutters will do a variety of boring operations, and
boring-bar components are standard items for the constructions of special bars. The
tool designer should investigate and seriously consider the use of commercial items
before designing a special bar which will require several hours of a skilled toolmak-
er’s time to manufacture. A few of the standard commercial boring-bar components
are mentioned in the following discussion. Their utilisation is limited only by the tool
designer’s imagination.
The expanding block-type boring tool (Fig. 4.104) is fitted into a slot in the bar and
held in position by a pointed screw. The insert blades can quickly interchanged, allow-
ing utilisation of various cutting-tool materials. The cutter support screws take the
304 Tool Design

cutting thrust at an angle, as shown, thus preventing the screws from working loose.
The block is provided with centreing holes to facilitate cutter grinding without arbors.
The blocks can be positioned in bars so that either forward or boring is possible. They
are also easily interchanged and preset to size in the tool room.

Fig. 4.104 The expanding block-type boring tool

Block boring tools minimise both bar deflec-


tion and tool wear because two opposing
cutters divide the load. Feed rates may
also be twice as fast as those of single-
point tools for the same reason. They are
readily available with single cutters, dou-
ble cutters, multiple cutters, micrometre-
adjustable cutters, and throwaway carbide
insert cutters. They can be used on either
stub or line boring bars.
Many boring-tool manufacturers now have
micrometre-adjustable cartridge-type
tools for use in commercial boring bars
designed by the user. Figure 4.105 shows Fig. 4.105 Adjustment of screw-on
the adjustment of a screw-on boring car- boring cartridge (Devlieg
tridge. This type of cartridge can be pur- Microbore)
chased as a standard item with a variety of
different cutters and tool geometries, including carbide throwaway inserts.
A commercial clamp-on type cartridge suitable for the construction of special bars is
shown in Fig. 4.106, and its application is shown in Fig. 4.107.

Special bar tools Whenever production justifies the expenditure, special bar tools
for boring, reaming, recessing, grooving, undercutting, and similar operations may be
designed. Simple bars may be designed entirely by the company’s tool designers,
but where the jobs are complex and likely to involve time-consuming problems, it is
generally more economical to design the bars in co-operation with the engineers of
companies specialising in this sort of work.
Design of Cutting Tools 305

Fig. 4.106 Commercial clamp-on boring Fig. 4.107 Application of clamp-on boring
cartridge (VR/Wesson Company) cartridge (VR/Wesson Company)

As an illustration, the bar shown in Fig. 4.108 was designed by Skully-Jones and
Company. Figure 4.108 shows a tool for chamfering both ends of a cylinder bore. This
is mounted in a special machine similar to a boring mill. The workpiece is held in a
fixture, and the tool is hand-fed into it with a linear motion until the stop collar touches
the pilot bushing. The tool, on continuing forward against the pressure of the spring is
activated through the central shaft. On the front of this is a wedge, which in turn forces
the end chamfering cutter against the workpiece. At the same time, the fixed cutter
touches the workpiece. When the cutting operation is finished, the spring retracts the
control shaft, and this in turn catches the toggle level, be withdrawing the end cutter
into a retracted position. This permits the bar to be withdrawn.

Boring-tool geometry Tool geometry for boring is much more critical than for turn-
ing because of the nature of the operation. The curvature of the bored hole and the

End chamfering Workpiece Chamfering Driving key Pilot bushing Draw key
cutter cutter

Jig plate Ball thrust bearing

Cutter adjustment
Pivot Main return spring

Fig. 4.108 A custom-designed chamfering tool


306 Tool Design

problem of chip disposal, coupled with the inherent lack of rigidity of a boring bar, all
combine to demand a more exacting tool geometry. The chip must flow away from
the finished surface and leave the bore in a manner that will not mark the surface.
This is also true for a finished turned surface, but here a great deal more open space
is available. The true rake of a boring tool becomes important, in that it must direct
the chip flow toward the centre of the bore and away from the finished surface. Since
the side-rake angle has the greatest influence on the direction of chip flow toward the
bore centre, it is generally the variable that is increased to give the proper true rake
for boring operations. Generally, the side-rake angles suitable for single-point turning
are satisfactory for the majority of boring operations where chip interference with the
finished surface is not a problem. When chip interference occurs, the side rake must
be modified to direct the chip flow toward the centre of the bore (when referring to bor-
ing tools, side rake is often referred to as axial rake and back as radial rake).
Remember that tool signature for boring tools is specified according to the way the
tool is held in relation to the workpiece, as explained earlier in this chapter. For exam-
ple, the side cutting-edge angle will be
determined by the position of the tool
in the bar. A tool held at 45° and cut- 45°
ting to a square shoulder must have a
side cutting-edge of 45°, although the
tool signature for the boring operation
would specify a 0° side cutting-edge
angle (see Fig. 4.109a).
A slight lead angle is recommended (a)
when boring through hopes or where a
40°
90° shoulder is not required, as shown
in Fig. 4.109b. A lead angle of 5 to 15° 30°
may increase cutter stability and has a
tendency to thin the chip. It also helps
to prevent the chipping of castings as
the cutter leaves the bore. 5°
The boring-tool nose radius, com-
bined with the rake of feed, determines
the surface finish produced by a bor- (b)
ing tool. Surface finish is often critical
Fig. 4.109 Side cutting-edge angles for 90°
operations, in that subsequent finishing
shoulders and through holes
operations may not follow boring.
Boring tools are often designed to cut on the recent line of the workpiece, as shown
in Fig. 4.110, especially for general boring operations of the engine lathe and other
general-purpose machine tools where vertical adjustment of the boring of the boring
bar is easily accomplished. In some instances, however, it may be desirable to set
the boring bar above the centreline of the work piece. For example, raising the tip of
the boring tool above the centreline the front relief and clearance angles of the tool
in relation to the workpiece. Additional relief and clearance could be attained when
machining small bores where the concave bore surface causes interference. By the
same token, raising the boring-tool tip above centre to increase relief also decreases
the back-rake angle. More back rake is required to attain the same “on-centre” cut-
ting action. In other words, when the top of the tool is above the centreline of the
Design of Cutting Tools 307

work-piece, it is necessary to grind back rake


on the tool to give it a true cutting rake of 0°
(see Fig. 4.111). The amount of back rake to
use for different tool settings varies according
to bore diameter. Table 4.18 shows the nec-
essary back rake for varying conditions. The
neutral-angle formula should be used when
the bore exceeds 6 mm.
Table 4.18 also shows the end-relief angles
for varying conditions. The angles shown
are those which must be ground to obtain a
proper working or bore relief when the tool
is set at a specified height above the cen- Fig. 4.110 Boring tool on centre
treline in a bore of a given size. Note that the line of workpiece
required relief angle becomes smaller as the
tool is raised above the centreline. The values
given pertain to the end-relief angles normal
to the bore. When the tool is set at 90° to the
bore, the end-relief angle of the tool becomes
the relief angle normal bore. However, when
the tool is set at an angle to the axis of the
bore, the normal bore relief angle becomes
less than the end relief angle of the tool (see
Fig. 4.112). It is therefore necessary to note
the position of the tool when specifying end-
relief angles for boring toots. When grinding
the tool, it should be held in the same relation
to the grinding wheel as it is to the bore. Fig. 4.111 Positive back rake for on-
centre cutting action
4.8 MILLING CUTTERS
The typical tool designer of today does not have to be concerned with the design of
milling cutters, a field best served by the specialists who work for milling-cutter manu-
facturers. The typical tool designer is now involved with selecting the best cutters for
a particular job, along with specifying sharpening angles for the various workpiece
materials to be machined. The majority of the cutters selected will be from a group
referred to as standard cutters. Standard milling cutters are generally recognised as
those which meet the requirements of the Metal Cutting Tool Institute, as specified
in the “Metal Cutting Tool Handbook”. The great majority of milling operations can be
performed with standard milling cutters.
When a special milling cutter is required, it is suggested that the tool designed consult
one of the major milling-cutter manufacturers. They are equipped to design and pro-
vide a special cutter and the milling process. It should also give him an understanding
of the elements of milling-cutter design.

The milling process The milling process, sometimes described as removing metal
advancing a work-piece against a rotating multiple-point tool, is more complicated
that the turning process. Cutting forces, angles of entry, and effective geometry of
tools metal change during the milling process because the cutter tooth is constantly
changing positions in relation to the work-piece.
Table 4.18 Rake and relief angkes for precision-boring tools set above centre
Height above CL Bore Height
above CL
Neutral angle A
308 Tool Design

CL work
CL work
C
Radius of bore
Clear bore by 1
32

Height above centre line, mm End relief (bore relief) angle C, deg
Bore 0.25 0.4 0.8 1.16 2.4 3.2 A=0 A = 0.25 A = 0.8 A = 1.6
diam, Neutral angle, deg* Group Group Group Group Group Group Group Group
mm I† II‡ I† II‡ I† II‡ I† II‡

1
8 4 6 11 __ ... ... ... 10 13 9 12 3 5
2
1 1
9.5 3 __ 5 9 __ ... ... ... 10 13 9 12 4 6
2 2
11 3 4 8 ... ... ... 10 13 9 12 5 7
1 1
12.5 2 __ 3 __ 7 ... ... ... 10 13 9 12 6 8
2 2
1 1
16 2 3 5 __ 11 __ ... ... 10 13 9 12 6 9 2 4
2 2
1 1 1
19 1 __ 2 __ 5 9 __ ... ... 10 13 9 12 6 9 2 5
2 2 2
1 1__ 1
22 1 2 3 __ 7 12 __ ... 10 13 9 12 6 9 2 5
2 2 2
Height above centre line, mm End relief (bore relief) angle C, deg
Bore 0.25 0.4 0.8 1.16 2.4 3.2 A=0 A = 0.25 A = 0.8 A = 1.6
diam, Neutral angle, deg* Group Group Group Group Group Group Group Group
mm I† II‡ I† II‡ I† II‡ I† II‡

1
25 1 2 3 __ 7 11 ... 10 12 8 11 6 9 2 5
2
1 1
32 ... 1 __ 3 6 8 __ 11 9 12 8 11 6 9 3 6
2 2
1 1
38 ... 1 2 __ 5 7 9 __ 9 12 7 10 6 9 3 6
2 2
45 ... 1 2 4 6 8 8 11 7 10 6 9 4 7
1 1
51 ... 1 2 3 __ 5 __ 7 8 11 7 10 6 9 4 7
2 2
1 1__
64 ... 1 1 __ 3 4 6 7 10 6 9 6 9 4 7
2 2
1 1
76 ... ... 1 2 __ 3 __ 5 7 10 6 9 6 9 5 8
2 2
89 ... ... 1 2 3 4 7 10 6 9 6 9 5 8
1
102 ... ... 1 2 3 3 __ 6 8 6 9 6 9 5 8
2
1
127 ... ... 1 1 __ 2 3 6 8 6 9 6 9 5 8
2
1
153 ... ... ... 1 2 2 __ 6 8 5 8 6 9 5 8
2
* To nearest 1/2°.
** Group I materials: bronze, cast iron, steel, malleable iron, semisteel, steel.
** Group II materials: aluminium, copper, fiber, magnesium, plastics, rubber, zine alloy.
Design of Cutting Tools 309

SOURCE: R L. Grand (ed.). “The New American Machinists’ Handbook,” McGraw-Hill, New York, 1955, by permission.
310 Tool Design

Bore relief (end-relief angle


50° normal to bore) is reduced

8.5°

45°
End-relief angle
of cutting tool
12°

Fig. 4.112 End relief normal to bore reduced when tool is set at angle in bar

The milling process is generally divided into two basic forms, referred to peripheral
milling and face milling. In peripheral milling, the finished surface is parallel to the axis
of the milling cutter and is generated by teeth located on the periphery of the cutter.
For face milling, the finished surface is at a right angle with the cutter axis and is gen-
erated by teeth located on the periphery and the flat end of the cutter. Figure 4.113
shows the position of the cutter in relation to the finished surface for both peripheral
and face milling.

Cutter axis

Cutter axis

Finished Finished
surface surface

(a) (b)

Fig. 4.113 (a) Peripheral and (b) face milling

In peripheral milling, metal removal may be accomplished by rotating the cutter against
the direction of workpiece traval, known as conventional or up milling, or by rotating
the cutter with the direction of workpiece travel, known as climb or down milling. At
first glance, it would appear that there would be difference in climb and conventional
milling, but a closer investigation reveals that the difference is considerable.
First of all, there is a difference in the formation of the chip, as shown in Fig. 4.114.
The chip being formed by conventional milling practise will be thin at the beginning
Design of Cutting Tools 311

and increase to a maximum at the end of the cut. The reverse will be the case for a
chip formed by climb milling.

Cutting
forces

Cutting
forces
Feed Feed

(a) (b)

Fig. 4.114 (a) Conventional (up) and (b) climb (down) milling

With conventional milling, the chips are formed by the cutter entering the already
machined surface at a very shallow angle. The cutter tooth tends to slide along the
surface at a short distance until enough pressure is built up to break through. This
sliding under pressure tends to dull the cutter, and the alternate sliding and sud-
den breaking through leaves exaggerated revolution marks on the finished milling
surface. This action becomes more pronounced with milling heat-treated alloys and
work-hardening materials because the tooth must cut through previously machined
material.
Chips formed by climb milling are formed by the cutter tooth entering virgin metal at
a much steeper angle. Here, full engagement of the tooth with the work is practically
instantaneous. This prevents a gradual building up of peripheral pressures and the
resulting sliding and dulling of the cutter mentioned above. It further allows the tooth
to leave the work-piece gradually, so that feed marks are not exaggerated.
Climb milling is especially desirable with milling cutters having a high radial rake
angle. The high radial rake angle provides a sharper cutting edge that penetrates
the workpiece easier but at the same time weakens the tooth because it is relatively
thin. The forces imposed on the tooth during conventional milling are in a direction
that will cause the tooth to deflect because it has no support to oppose the forces. On
the other hand, the forces imposed on the tooth during climb milling are in a direction
more parallel to the tooth body, thus giving added support to prevent deflection (see
Fig. 4.115). Cutters designed for high metal removal rates and high cutting efficiency
through the use of high positive radial rake angles should be used under climb-milling
conditions whenever possible.
Climb milling forces the workpiece against mill table to give the workpiece more sup-
port especially when milling thin workpieces. Narrow metal-slitting saws do not “walk”
or cut crooked slots under climb-milling conditions. In fact, the only instance where
climb milling is not advantageous is when the workpiece has a hard scale or surface
inclusions that would quickly dull the cutter teeth. In this case, it is better to approach
the surface from the underside for longer cutter life.
312 Tool Design

Fig. 4.115 Climb-milling and high-radial-rake milling cutters

This discussion of climb and conventional milling seems to indicate that conventional
milling should generally not be considered. On the other hand, older machinists will
tell you that there was a time when climb milling was an exception rather than the
rule. All milling was done conventionally. This was (and still is) due to the condition
and design of the milling machine. Feed movements on older milling machines utilise
a lead-screw arrangement that allows backlash (play between the screw and nut),
especially when the lead screw becomes worn. Backlash is not a problem when mill-
ing by the conventional method, in that the force moving the cutter and the one moving
the table oppose each other, as shown in Fig. 4.114. The backlash in the lead screw
is always kept in one direction. However, in climb milling, the forces are nearly in the
same direction and the backlash may be “jerked” in the opposite direction when the
milling tooth engages the cutter. The result is an overload on the tooth, and either the
cutter is broken or a bad case of chatter develops. It is therefore absolutely essential
that the milling machine be designed correctly before attempting climb-milling opera-
tions. This is usually accomplished by providing the milling machine with a backlash
eliminator, which usually consists of two opposing nuts on the screw that automati-
cally take up backlash when the screw rotation is reversed. The machine itself should
be a fairly heavy type with all gibs and slides free from play and backlash.
Face milling approximates conditions found in both climb and conventional milling,
provided the cutter is carrying an equal load on each side of its own centreline. Figure
4.116 shows a chip that is thin at tooth entrance, increasing to maximum at the cen-
treline of the cutter and then decreasing to a minimum as the tooth leaves the work-
piece. The forces of climb milling which would jerk backlash out of the lead screw are
cancelled by those of conventional milling as long as an equal amount of metal is
being removed on each side of the cutter centreline.
Design of Cutting Tools 313

Climb-milling
effect

Conventional-
milling effect

Chip thickness at maximum

Fig. 4.116 Tooth pattern of face-milling operation

Referring again to Fig. 4.116, note that the cutter tooth picks up its chip load under
the conventional effect and therefore tends to slide. This action is detrimental to cut-
ter life, especially with carbide cutters, where the chip thickness should be kept fairly
constant. It is therefore important that the diameter of the face-milling cutter be wider
than the workpiece in order to increase the chip thickness at points of tooth entry and
exit. A good rule of thumb is to provide a cutter diameter-to-work ratio of approxi-
mately 1.5 to 1.

Classification of milling cutters There seems to be no clear method of classifying


milling cutters, due to the many different types available in the market today. Milling
cutters are often identified according to the type of construction, such as inserted-
tooth cutters, solid cutters, and carbide-tipped solid cutters. On the other hand, the
same cutters may be referred to by the method of mounting, arbor-type, shank-type,
and spindle-mounted milling cutters being typical examples. To go one step further,
some milling cutters are classified according to their use or application, such as T-slot
cutters and Woodruff key-seat cutters. T-slot and Woodruff cutters are both solid cut-
ters of the shank type.
The various methods of classification are quite clear to the experienced tool designer
or machinist. He understands that reference to one type of cutter depends upon cir-
cumstances. He may refer later to the same cutter by a different classification. This
practise is often confusing to the beginner, and the only way to understand the milling-
cutter family clearly is through experience.
However, there is one method of classifying milling cutters which is not confusing and
which applies to all milling cutters with few exceptions. There are only two methods of
providing relief on milling cutters, and therefore a system of classification according to
relief is a basic classification that causes no confusion.
The two basic cutter types according to relief are the form- and profile-relieved cut-
ters. The profile- relieved cutter is obtained by sharpening a narrow land behind the
cutting edge. The land is retargeted when the cutter becomes dull. Form-relieved cut-
ters have a curved relief behind the cutting edge and are sharpened by grinding the
face of the tooth. The shape and resharpening methods of a form- and profile-relieved
tooth are shown in Fig. 4.117.
The shape of profile-relieved cutters is determined by the sharpening operation. The
form produced by profile-relieved cutters is usually a plane surface because it is eas-
ier to sharpen a tooth shape which will produce a plane surface. For example, it is
much easier to pass a cutter tooth along a grinding wheel in a straight line than in a
314 Tool Design

curved or irregular line. Form cutters are generally not profile-relieved for this reason.
This does not mean that form cutters cannot be profile-relieved. Form cutters can be
sharpened on the periphery by using special attachments and sharpening machines.
It is sometimes necessary to use profile-sharpened form cutters when machining
small and fragile workpieces to take advantage of their increased efficiency. A profile-
sharpened cutter will be more free-cutting than an equivalent form relieved design.

Grinding
wheel

Curvature of tooth Relief


forms relief

(a) (b)

Fig. 4.117 Methods of resharpening milling cutters. (a) Form relief. Relief is provided at the
time of manufacture; the face of the tooth is ground on a radial line to resharpen
(b) Profile relief. Relief is ground behind the cutting edge in the form of a land

It should be noted that retargeting the side of a profile-relieved tooth reduces its width.
Interlocking two or more milling-cutter sections is often used to provide adjustment
or compensation for size after retargeting. By using spacers of various thickness, the
width of the cut can be varied. Figure 4.118 shows typical interlocking cutters.

Fig. 4.118 Interlocking milling cutters (Barber-Colman Company)

The relief of a form-relieved milling cutter is provided at the time of manufactures. The
form is created by a form tool, and sharpening is accomplished by grinding the face
Design of Cutting Tools 315

of the tool on a radial line. The main advantage is the case of resharpening a cutter
whose from contains intricate shapes that would be difficult to follow by grinding of
the profile. Normally, a from-relieved cutter will not change shape during the life of the
cutter, providing it is ground on a radial line.
Form-relieved cutters ground on a radial line will have zero rake. Some cutters are
designed and manufactured with positive rake. In sharpening a cutter of this type,
care must be taken to offset the cutter centreline in order to maintain the original rake
angle. The amount of this offset, measured on a horizontal plane, is usually stamped
on the cutter.
The use of form relief on milling cutters is not new. It was introduced over 100 years
ago by Joseph R Brown when he introduced a method of producing a cutter of practi-
cally any profile which could be retargeted without changing its shape. Form relief is
produced on a special lathe called a backing-off lathe. The lathe tool, ground to the
shape to be reproduced in the cutter, is advanced by a cam as the cutter rotates to
produce the desired relief (see Fig. 4.119). At the end of the relief (heel of tooth) the
tool holder and tool snap back immediately so that the form is ready to start cutting
again at the front of the next cutter tooth. The cam action must be timed to correspond
to the number of teeth in the cutter. The drop of the cam is determined by the amount
of relief and the diameter and number of teeth in the cutter. The cam is determined by
the amount of relief and the diameter and number of teeth in the cutter. Most formed
cutters have a relief angle of approximately 12°. The amount of cam drop to produce
a 12° relief is determined by

tan 12° × cutter OD


Cam drop = __________________
number of cutter teeth

Cutter is sharpened on
face of tooth only
Cam drop Form tool
(backing off)

Cross slide

Cam action timed to correspond to


number of teeth in cutter

Fig. 4.119 The principle of the backing-off lathe to produce form relief
316 Tool Design

Profile relief is more frequently used and is found on end mills, plain and side mills,
saws, face and shell mills, angle cutters, etc. Profile-sharpened form cutters, because
of the difficulties involved in resharpening, are generally used only for simple forms.
Form relief is used on gear-tooth cutters, concave and convex cutters, corner-round-
ing cutters, and special cutters involving complicated forms. Form-relieved milling
cutters can be designed and manufactured to mill almost any form if the part can be
positioned and held so that the entire form is accessible to the cutter. Figure 4.120
shows several examples of standard milling cutters. The following standard-cutter list-
ing gives the name of the cutter, general description, and type of relief:

(a)

(b)

(c)

(d) (e) (f)

(g) (h) (i)

Fig. 4.120 Standard milling cutters: (a) Double-angle (b) Woodruff key (c) Concave
and convex (d) Plain (e) Side (f) Staggered-tooth side (g) Interlocking side
(h) Metal-slitting saw (i) Screw-slotting (Pratt & Whitney)
Design of Cutting Tools 317

Plain milling cutter (profile relief) A cutter of plain cylindrical form having teeth on
the circumferential surface only. Teeth may be either straight or helical. This cutter is
used for slabbing.
Side milling cutter (profile relief) A cutter of cylindrical form having teeth on the cir-
cumferential surface and also on both sides. These cutters are frequently used in pairs
for milling both ends of the work at the same time. This is called straddle milling.
Half-side milling cutter (profile relief) A cutter similar to a side milling cutter but with
teeth on one side only. In both side and half-side cutter, the side teeth extend a por-
tion of the distance from the circumference toward the axis.
Interlocking-side milling cutter (profile relief) Similar to a side milling cutter except
that it is composed of two interlocking section that can be adjusted by the use of shims
to form a complete cutter of exact width.
Staggered-tooth milling cutter (profile relief) A cutter of cylindrical form having cut-
ting teeth on the circumferential surface only, the teeth being alternately of opposite
helix or angle. This type of cutter has side teeth extending a short distance toward the
axis for the purpose of chip clearance only; these side teeth are not ground for cutting
purposes.
Single-angle milling cutter (profile relief) A cutter having teeth on the conical surface
and with or without teeth on one or both of the flat sides. Designated by the included
angle between the conical face and the larger flat surface.
Double-angle milling cutter (profile relief) A cutter having two intersecting conical
surfaces with teeth on both. The angle of teeth may not be symmetrical with respect
to a plane at right angles to the axis.
Metal-slitting saw (profile relief) A plain milling cutter with sides relieved or “dished”
to afford side clearance. It has more teeth than a plain milling cutter and generally has
a thickness less than 3/16 mm.
Screw-slotting cutter (profile relief) A cutter used for shallow cuts as in screwdriver
slits. It has fine teeth on its circumference and is not ground on the sides.
End mill (profile relief) A cutter with teeth on the circumferential surface and one
end. The shank may be straight or tapered. The teeth may be helical or parallel to the
axis of rotation. A spiral end mill is an end mill with a moderate helix angle.
Shell end mill (profile relief) Similar to an end mill with a hole for an arbor instead
of having a shank. Generally, driven from a key slot across the back of the face. The
teeth may be straight or spiral.
T-slot cutter (profile relief) Similar to an end mill but designed for cutting T slots.
Woodruff key-seat cutter (profile relief) Either a shank or arbor cutter designed to
cut semi-cylindrical keyways in shafts for Woodruff keys.
Gear cutter (form relief) A formed cutter specially designed to cut one space at a
time in gears.
Multiple gear cutter (form relief) A single cutter unit made to mill several gear spaces
at one pass. Generally used for roughing out gear blanks.
318 Tool Design

Convex cutter (form relief) A formed cutter designed to mill a concave surface equal
to a half-circle or less. Size is designated by specifying the diameter of the circular
form.
Concave cutter (form relief) A formed cutter shaped to mill a convex surface of cir-
cular contour equal to half a circle or less. Size is designated by the diameter of the
circular form.

Nomenclature of milling-cutter elements The Metal Cutting Tool Institute lists the
following as the important milling-cutter elements:
Angular flute A space between two cutter teeth which forms a cutting edge lying in
a plane intersecting the tool axis at an angle. It is unlike a helical flute in that it forms
a cutting edge which lies in a single plane (see Fig. 4.121a).
Axial run out The total variation in an axial direction of a cutter element from a true
plane of rotation.
Body The carrier or head for holding blades, or that part of a solid or tipped cutter
exclusive of the teeth or shank.
Chamfer (1) A bevelled surface to eliminate an otherwise sharp corner; (2) a relieved
angular cutting edge at a tooth corner.
Clearance The additional space provided behind the relieved land of a cutter tooth
to eliminate undesirable contact between the cutter and wordpiece (see Fig. 4.121b)
Corner angle On face-milling cutters, the angle between an angular cutting edge of
a cutter tooth and the axis of the cutter, measured by rotation an axial plane (see Fig.
4.121c).
Entrance angle The angle formed between a centreline on the cutter which is per-
pendicular to the direction of feed and a radial line through a point on the cutting edge
where the tooth first contacts the workpiece (see 4.121d).
Face cutting edge That edge of the tooth on a face mill which travels in a plane
perpendicular to the axis. It is the edge which sweeps the milled surface in the normal
operation of a face-milling cutter (see 4.121c).
Face cutting edge angle The angle of concavity between the face cutting edge and
the face plane of a face mill. It serves as relief to prevent the face cutting edges from
rubbing in the cut (see Fig. 4.121c).
Fillet The bottom surface of the flute (see Fig. 4.121b).
Flute The chip space between the back of one tooth and the face of the folling tooth
(see Fig. 4.121b).
Form cutter Any cutter, profile-sharpened or cam-relieved, shaped to produce a
specified form on the work.
Form tool As related to milling cutters, a tool used to shape a cutter blank or to pro-
duce the form on a cam-relieved cutter.
Heel (1) The back edge of the relieved land (see Fig. 4.121e); (2) the inner end of a
face-cutting edge (see Fig. 4.121c).
Design of Cutting Tools 319

Helical A term describing a cutting edge or flute which progresses uniformly around
a cylindrical surface in an axial direction.
Helix angle The cutting-edge angle which a helical cutting edge makes with a plane
containing the axis of a cylindrical cutter (see Fig. 4.121f).
Interlocking Mating cutter sections on which side projections on one section mesh
with those on an adjacent section to provide cutting-edge overlap (see Fig. 4.121g).
K land A relatively narrow land on the face of a tooth from the cutting edge inward
which is at a lesser rake angle than the main face of the tooth. It is the surface between
the chip and the whole tooth face (see Fig. 4.121d).
Land The narrow surface of a profile-sharpened cutter tooth immediately behind the
cutting edge (see Fig. 4.121h).
Cylindrical land A narrow portion of the peripheral land, adjacent to the cutting edge,
having no radial relief.
Relieved The portion of the land adjacent to the cutting edge which provides a
relief.
Lead (1) The axial advance of a helical cutting edge in one turn around the axis;
(2) the relieved angular cutting edge between the corner angle and the face cutting
edge of a face mill (see Fig. 4.121c).
Nesting The axial overlapping of cutters of different diameters without the use of
interlocking projections (see Fig. 4.21g).
Radial run out The total variation in a radial direction of all cutting edges in a plane
of rotation.
Rake The angular relationship between the tooth face or a tangent to the tooth face
at a given point and a given reference plane or line.
Axial rake Applies to angular (not helical) flutes. The axial rake at a given point on
the face of the flute is the angle between the tool axis and a tangent plane at the given
point (see Fig. 4.121c).
Effective rake The complement of the angle between the direction of motion of any
point on the cutting edge and the direction of chip flow from the same point. Effective
rake is therefore the rake resulting from three factors: (1) the cutter geometry, (2) the
actual path of the cutting edge, and (3) the actual direction of the chip flow.
Helical rake Applies to helical teeth (not angular). The helical rake at a given point
on the flute face is the angle between the tool axis and a tangent plane at the given
point (see Fig. 4.121f).
Hook A concave condition of a tooth face. The rake of a hooked tooth face must be
determined at a given point (see Fig. 4.121i).
Negative rake Describes a tooth face in rotation whose cutting edge lags the surface
of a tooth face (see Fig. 4.121d).
Positive rake Describes a tooth face in rotation whose cutting edge leads the surface
of the tooth face (see Fig. 4.121j).
320 Tool Design

Radial rake The angle between the tooth face and a radial line passing through the
cutting edge in a plane perpendicular to the cutter axis (see Fig. 4.121j).
True rake See Effective rake.
Relief The result of the removal of tooth material behind or adjacent to the cutting
edge to provide clearance and prevent rubbing.
Design of Cutting Tools 321

Negative radial rake

Entrance Face ridge


angle
K land

Workpiece

(d)

Cutter sweep Neck Heel


Flute
length

End Conventional Raised


gash land land
Helix angle

(e)

Face width
Helical teeth
Helical rake
angle, left-hand
helix shown

(f)

Nested Interlocked
1 2 2 1

(g)
322 Tool Design

Primary relief
Relieved Offset
Cylindrical
land land
90°

CL CL CL

(h)

Tangential
rake

Hook

(i)
Radial rake angle
positive shown
Radial relief Tooth face
Wear land
Tooth
Flute Axial relief
Fillet

Offset
(l)
(j)

Radial
Internal External relief
back taper back taper Relief
angle
Axial
relief
angle

Cam relief
Concavity
Flat Flat
relief relief

(k)

Fig. 4.121 Milling-cutter elements; (a) Angular flute (b) Peripheral milling cutter (c) Face-
milling cutter (d) Face milling (e) End-milling cutter (f) Plain milling cutter
(g) Nested and interlocking cutters (h) Types of lands (i) Hook of cutter tooth
(j) Form-relieved cutter (k) Types of relief (l) Wear land (The Metal Cutting
Institute)
Design of Cutting Tools 323

Axial relief The relief measured in the axial direction between a plane perpendicular
to the axis and the relieved surface. It can be measured by the amount of indicator
drop at a given radius in a given amount of angular rotation.
Back taper A slight reduction of the outside diameter from front to back of the essen-
tially cylindrical in which the cutting edges lie (see Fig. 4.121k).
Cam relief The relief from the cutting edge to the back of the tooth produced by a
cam-activated cutting tool or grinding wheel on a relieving (back-off) machine (see
Fig. 4.121k).
Eccentric relief A convex relieved surface behind the cutting edge (see Fig.
4.121k).
Primary relief The relief immediately behind the cutting edge.
Relief angle The angle formed between a relieved surface and a given plane tangent
to a cutting edge or to a point on a cutting edge (see Fig. 4.121k).
Radial relief Relief in a radial direction measured in the plane of rotation. It can
be measured by the amount of indicator drop at a given radius in a given amount of
angular rotation (see Fig. 4.121k).
Secondary relief See Clearance, the preferred term.
Shank The projecting portion of a cutter which locates and drives the cutter from the
machine spindle or adapter.
Taper shank A cutter shank made to fit a specified (conical) taper socket.
Tooth face The surface of the tooth on which the chip impinges (see Fig. 4.121b).
Wear land The cylindrical or flat land worn on the relieved portion of the cutter tooth
(see Fig. 4.121l).

Selection of cutter geometry and design The major elements to consider when
selecting a milling cutter for a particular application are size, teeth, flutes, relief, and
material. The following discussion gives a brief general description of these elements
and how they may be influenced by operating conditions:
Size The face width of the cutter must be wide enough to provide adequate support
behind the cutting edges. The cutter diameter should be kept as small as possible
because small diameter cutters require less torque and deflect less. The diameter will
depend upon the depth of the flute and the diameter of the hole in the cutter. A gen-
eral rule for diameter is to provide a minimum ratio 3:1 between the cutter diameter
the and hole diameter. Larger diameter cutters may be required when interference
between the outer arbor support and workpiece occurs.
Teeth The solution of the correct number of teeth in a milling cutter is a compro-
mise depending upon working material, type of milling cutter, and the surface fin-
ish required. A milling cutter with a relatively large number of small teeth around its
periphery will promote smoother cutting because more teeth are in the cut. When
other cutting conditions are constant, a cutter of many teeth will also have a finer feed
per tooth, which results in a better surface finish. However, a cutter of may teeth may
not have adequate chip space to prevent chip interference. A soft material that allows
a heavy feed rate and produces long continuous chips requires a greater chip space.
324 Tool Design

An effective way of providing chip space is by reducing the number of teeth in the
cutter and increasing the size of the individual tooth. This provides a stronger tooth
and is therefore suitable for heavy roughing cuts when a large volume of material is
being removed.
A milling cutter should have enough teeth to ensure uninterrupted contact with the
workpiece and yet not so many that there is too little space between the teeth for
chip disposal. A rule of thumb is to select a cutter so that a minimum of one tooth will
always be in the cut. This will ensure that one tooth will not leave workpiece before
another enters and will promote a smooth cutting operation. Additional teeth may be
added to improve the surface finish, but the trend is to specify a cutter having a coarse
tooth when heavy and wide cuts are necessary. In this instance, it has been found that
coarse-tooth cutters with fewer teeth will stand up longer than fine-tooth cutters with
many teeth. This does not hold true when using narrow plain mills and slitting saws.
(This main advantage of specifying a fine-tooth milling cutter is that a smoother sur-
face finish is produced without reducing the feed rate.)
Regarding the number of teeth and type of cutter, fewer teeth can be used on a face
mill because the tooth contact is usually longer. In other words, each tooth is in the cut
for a longer period of time.
Milling cutters designed for machining aluminium alloys require additional chip space,
which in this instance is often provided by increasing the size of the cutter flute. This
weakens the tooth but is acceptable because milling soft aluminium does not require
maximum tooth strength.
Flutes The flutes of a milling cutter can be straight, helical or angular. Helical flutes
are used on the majority of cutters designed for wide peripheral cuts. Helical teeth
form the chip at an angle with respect to the direction of feed. This provides a shear-
ing action, and the helix angle of the flute is sometimes incorrectly to as the “shear
angle” of the cutter.
When milling-cutting teeth are cut on a helix, the entire cutting edge does not come
into contact with the workpiece at one time. The helix cause more than one tooth
to be in contact with the workpiece along a single contact line parallel to the cutter
axis. Each individual tooth picks up the chip gradually until a maximum chip load is
attained. The maximum clip load continues for a considerable part of the tooth travel
and then decreases gradually as the tooth leaves the workpiece. This action provides
smooth and continuous cutting and greatly decreases the tendency to chatter.
The helix may be either right or left hand. To determine whether the flute helix is right
or left hand, visualise it as a common screw thread (another form of helix). If is a
right-hand thread, it is a right-hand helix and vice versa. Standard end mills are made
in various combinations of hands of rotation and hands of helix. A right-hand cut and
right-hand helix is excellent for milling slots and shoulders, in that chips are augered
away from the finished work surface. However, there is a tendency for them to suck
in or grab with some types of work materials because the tooth tends to act like a
thread. A right-hand cut, left-hand helix would help offset this condition because the
helix would force the cutter in the opposite axial direction. Right-hand cut – left hand
helix end mills are also well suited for profile operations when the cutter is used to cut
on its periphery only. They tend to push the workpiece down toward the heavy part of
the machine and provide a much steadier cutting action.
Design of Cutting Tools 325

The cutting action of a tooth formed by a straight flute is intermittent. When the tooth
enters the workpiece, the whole length of the tooth takes the full load and the cutting
forces increase rapidly. The forces continue to increase until the tooth leaves the cut
and then suddenly drop. The shock load produced by this type of cutting generally
reacts through the drive mechanism of the machine to produce chatter.
The majority of form relieved cutters have straight flutes because of the difficulty in
sharpening form relieved cutters. Narrow plain milling cutters and slitting saws may
have straight flutes. In general, however, the use of milling cutters with straight flutes
should be avoided.
Angular flutes on milling cutters are a compromise between helical and straight flutes.
They are best suited to cutters with symmetrical forms and narrow face widths. The
face of the tooth is at an angle to the cutter axis and therefore the radial rake varies
from one side of the cutter to the other. They are not recommended for wide cutters
for this reason. Special sharpening equipment is not needed for angular flute cutters.
It is only necessary to align the swivel table of the cutter grinder with the angle of the
tooth and sharpen through the gash in a straight line.
Relief Relief has already been discussed in general terms. In review, two methods
of relief are provided which actually define two basic cutter types. These cutter types
are the form and profile relived cutters, and the names are descriptive of their design.
The following discussion will be concerned with the specifics of relief.
The relief is not affected when resharpening form-relieved cutters because only
the faces of the teeth are ground. The relief is built in at the time of manufacture.
Resharpening profile relieved cutters consists of removing enough from the top of
the teeth back of the cutting edge to remove the wear land. The angle of relief to be
sharpened into the cutter depends upon the diameter of the cutter, the cutting tool
material, and the material to be milled. The angle must be greater for small cutters
than for large ones in order to prevent heel drag.
Relief angles should be just enough to eliminate heel drag. This practise leaves
more metal for heat dissipation and ensures maximum strength at the cutting edge.
Excessive relief angles tend to increase the likelihood of chatter and may add to the
causes of failure during heavy milling operations.
The hardness of the material has an effect on the correct relief angle. Larger relief
angles may be used with softer materials because maximum strength at the cutting
edge is not required. Average relief angles for work materials of varying harness are
given in Table 4.19.

Table 4.19 Average relief angles for work materials of varying hardmess
Work material
Cutter Tool material Steel Cast Non-ferrous
iron and non-metallic

High-speed steel 5-10° 5-10° 7-12°


Peripheral or OD cutting edges* Cast alloy 4-6° 4-6° 5-10°
Cemented carbide 4-6° 4-6° 5-10°
Side or end cutting edges All 1-4° 1-4° 2-7°
* Smaller-diameter cutters require larger relief angles.
SOURCE: Barber Colman “Milling Cutter Handbook.”
326 Tool Design

The width of land (the narrow surface immediately behind the cutting edge that is
ground to the relief angle) will depend upon the type and size of the milling cutter.
The land must be narrow enough to prevent the heel from dragging on the finished
workpiece surface. This may vary from 0.13 to 0.25 mm on small end mills and up
to 3 mm on large diameter cutters, with 0.8 to 1.6 mm as an average. The land may
become too wide as a result of repeated sharpening, and the heel of the tooth will
drag on the workpiece. To control the width of the land, a clearance angle (sometimes
called secondary relief) of twice the relief angle (approximately) is ground as shown
in Fig. 4.122.

Relief on cutter
after repeated
Relief on new sharpenings; heel
cutter (primary) drag is result

Cutter tooth Additional secondary Cutter tooth


relief

Fig. 4.122 The width of land must be narrow enough to prevent heel dreg

The relief on side cutting edges may be less than that on peripheral cutting edges
because side cutting edges do not cope with a radial surface. Heel drag is not a prob-
lem because it does not exist. Relief on side cutting edges is usually one-quarter to
one-half that of the peripheral cutting edges.
The relief on end teeth of end and face mills may also be less for the same reason.
Multiple-tooth high-speed-steel end mills are generally ground with about 4° relief,
although two-flute end mills may have as much as 7° because they are frequently
plunged endwise into solid stock.
End and side teeth are often sharpened with radial taper by removing from 25 to 50
mm more material near the centre of the cutter than at the outside. This prevents the
teeth from rubbing on the finished workpiece surface.
It has been previously noted that relief must be increased as the diameter of the
milling cutter becomes smaller. The smaller cutters must have extremely large relief
angles in order to prevent heel drag. For example, 6 mm cutter has from 12 to 15°
relief, and a 3 mm cutter has from 16 to 19° relief. Table 4.20 gives relief angles for
the smaller cutters.
The necessary increase in relief on small cutters results in a decrease in cutting-edge
strength because of the removal of material behind the cutting edge. To increase the
strength of the cutting edge on smaller-diameter cutters, eccentric relief is often fab-
ricated into the cutting edge at the time of sharpening. Figure 4.123 shows the basic
difference between eccentric and regular flat relief. As can be seen in Fig. 4.123,
eccentric relief has the advantage of more solid metal backing up the actual cutting.
In milling and retargeting, this additional backing aids in dissipating heat. Sharpening
by the eccentric method will produce a smoother finish, and the cutting edges not be
honed after sharpening.
Design of Cutting Tools 327

Table 4.20 Relief angles for small-diameter cutters (General-purpose work)

Cutter diameter Primary relief, deg Primary land (mm) Secondary relief, deg
mm Min Max Min Max Min Max

1.59 22 26 0.25 32 36
2.38 18 23 0.25 28 32
3.18 16 19 0.25 0.38 28 31
3.97 15 18 0.25 0.51 26 29
4.77 14 17 0.25 0.51 26 29
5.56 13 15 0.25 0.51 24 26
6.36 12 15 0.38 0.51 22 25
7.15 12 15 0.38 0.51 21 24
7.94 12 15 0.38 0.64 21 24
8.74 11 14 0.38 0.64 19 22
9.53 11 13 0.38 0.64 19 21
10.33 11 13 0.38 0.64 19 21
11.12 11 13 0.51 0.76 19 21
11.92 11 13 0.51 0.76 19 21
12.71 10 13 0.51 0.76 18 21
14.30 10 13 0.51 0.76 18 21
15.89 10 13 0.64 0.89 18 21
17.48 9 12 0.64 0.89 18 21
19.07 9 11 0.76 1.02 18 20
20.65 9 11 0.76 1.02 18 20
22.24 9 11 0.76 1.02 17 19
23.83 8 11 0.89 1.27 16 19
25.42 8 10 0.89 1.27 16 18
28.60 8 10 0.89 1.27 16 18
31.78 8 10 1.02 1.53 14 16
34.95 7 10 1.02 1.53 13 16
38.13 7 10 1.02 1.53 13 16
41.31 7 9 1.02 1.53 13 15
44.49 7 9 1.02 1.53 12 14
47.66 7 9 1.02 1.53 12 14
50.84 6 9 1.02 1.53 12 14
57.20 5 8 1.02 1.53 12 14
63.55 5 8 1.02 1.53 12 14
69.91 5 8 1.02 1.53 12 14
76.26 5 8 1.02 1.53 12 14
328 Tool Design

Eccentric relief can be produced Additional metal provided


Eccentric relief
with standard cutter-grinding equ- by eccentric relief
ipment. The technique differs from
regular cutter sharpening in that
the eccentric is generated by the Flat relief
relative motion of the cutter tooth
and the grinding wheel. It is nec-
essary to dress the grinding-wheel
periphery at an angle or to tilt the
wheel at an angle with the cutter
axis. The angle depends upon the
helix angle of the cutter, because Fig. 4.123 The principle of eccentric relief
the helix determines the angle at
which the cutter tooth approaches the grinding wheel. Thus, a helical tooth is required
to generate eccentric relief, and this method cannot be used to produce eccentric
relief on straight flute cutters. The degree of relief is varied by changing the angle of
wheel inclination or the angle dressed on the wheel. The angle is determined by the
formula1
tan F = tan c tan h
where F = angle of wheel inclination or dress
c = relief angle of cutter tooth
h = helix angle of cutter
The measurement of relief angles on small-diameter cutter also poses a problem
because there is not enough diameter or tooth to use standard-cutter clearance
gauges. Eccentric relief cannot be measured accurately with standard-cutter clear-
ance gauges because there is no flat
land for comparison. A simple inexpen-
sive method for measuring relief angles,
often referred to as the indicator-drop
method, consists of two dial indicators
and bench centres. The cutter to be
maturated between the centres, and the
two indicators are positioned as shown
Indicator drop
in Fig. 4.124.

Geometry of face-milling cutters Face


milling differs from peripheral milling in
that the cutting edge is along the periph- Checking
ery and face of the cutter. The point where distance
the two cutting edges meet is the corner,
which may be sharply pointed, rounded Fig. 4.124 Measurement of relief angles
or chamfered to meet the requirements by the indicator-drop method
of the milling operation. A sharp square (From MCTI, “Metal Cutting
corner is generally avoided because of Tool Hand-book”)

1
More specific information on producing eccentric relief may be obtained from the Metal Cutting Tool Institute
or from the major manufactures of end-milling cutters.
Design of Cutting Tools 329

the tendency to chip off the fragile point, especially when using tungsten carbide as a
cutting-tool material. High-speed steel is stronger, but experienced machinists have
found that stoning a slight radius on the corner of a new high-speed-steel face or end
mill will reduce the tendency to chip.
When it is necessary to mill a square
shoulder, the sharp corner is often
eliminated by grinding a small chamfer
or radius of approximately 1.6 mm into
the corner of the cutter. In this case, the
radius or chamfer should not be too large
in order to allow an unrestricted flow of
metal. Metal flow is at approximately right
angles to the cutting edge. When a large Fig. 4.125 Formation of chip by large
chamfer or radius is used and the depth chamfer and depth of cut, which
of cut permits metal removal on both the permits metal removal on both
periphery and chamfer, the metal flow in periphery and chamfer
different directions and causes interfer-
ence within the chip. A large rounded corner has a similar result (see Fig. 4.125).
Chip formation is much more complex and distorted. Excessive heat is produced,
along with a reduction in the cutting efficiency. The end result may be excessive
chatter, poor chip formation, and shorter tool life. A small chamfer or nose radius
(1.6 mm or less) does not greatly interfere with the main flow of the chip. A short 45°
chamfer is often preferred because it can be sharpened into the cutter with minimum
effort. The sharpening of a true radius requires a special attachment for the tool and
cutter-grinding machine, and although the radius is not difficult to grind, the operation
is time consuming.
It is preferable to use a chamfer sufficiently wide for cutting to be confined to the cut-
ting edge along the chamfer when face milling a plane surface. The chamfer is formed
by the corner angle (the angle between an angular cutting edge of a cutter tooth and
the axis of the cutter, measured by rotation into an axial plane). The chamfer has the
effect of thinning the chip in the same manner that the chip is thinned by the side
cutting-edge angle of a single-point tool. The milling feed rate may be increased with-
out increasing chip thickness when the corner angle is increased. Corner angles of
15, 30, and 45° are commonly used, although a great deal of metal must be removed
when producing the initial 45° chamfer.
The geometry of a single tooth in a face-milling cutter is closely related to that of a sin-
gle-point cutting, as shown in Fig. 4.126. The side rake of a single-point tool becomes
the radial rake of the face-milling cutter, while the side cutting-edge angle becomes
the corner angle. The back rake of the single-point tool compares with the axial rake
of the face-milling cutter. The basic angles of the face-milling cutter combine to form
functional angles under cutting conditions which substantially determine the cutting
performance. The important angles are the true rake and angle of inclination.
The true rake of a face-milling cutter is the slope of the tooth face with respect to a
radial reference plane and is the resultant of the radial rake, the axial rake, and the
corner angle. It is in the plane at right angles to the path generated by the cutting
edge, as shown in Fig. 4.126.
330 Tool Design

Axial
True rake rake (+)
Inclination
Negative Reference (+)
plane Corner Reference
Axis angle plane

Face
Radial of
rake (–) tooth Axis

Radial
rake (–)

Fig. 4.126 The elements of face-milling cutter teeth, including the true rake and the angle
of inclination (Cincinnati Milacron, Inc)

The same component angles that determine the true-rake angle (the axial rake, the
radial rake, and the corner angle) also determine the angle of inclination of the cutting
edge with respect to the reference plane. The angle of inclination is sometime more
significant in tool performance than the true rake, because it allows, within limits, con-
trol of the direction of motion of the chip. A negative inclination directs the chip into the
cutter body, clogging the cutter and marring the machined surface, whereas a positive
inclination directs the chip away from the cutter.
A suitable selection of axial rake, radial rake, and corner angle makes it possible to
design a carbide face-milling cutter having a small negative true rake to provide satis-
factory tool life, together with a positive inclination for unrestricted chip flow, as shown
in the following example and Fig. 4.66.

Determine the true rake and angle of inclination for a face-mill-


Example 4.22 ing cutter with positive axial rake of 10°, negative radial rake
of – 30°, and 60° corner angle.
Solution: For the true rake, draw line T from the + 10° point on the axial-rake scale to
the black – 30° point on radial-rake scale. At its interdiction with the 60° vertical line
from the black corner-angle scale read the true rake of approximately 80°.
For the angle of inclination, draw line I from the + 10° point on axial-rake scale to the
grey – 30° point on the radial-rake scale. At its intersection with the 60° vertical line
from the grey corner-angle scale read and inclination of T = 30°.

Indexable-insert milling cutters The majority of indexable-insert applications for


milling cutters have been for face milling, although they have also been applied to
other types. Like the index able or throwaway mechanical single-point tool holder,
their main attraction is the elimination of the need to resharpen the cutting edge. All
that is required to obtain a new cutting edge is to index or replace each insert in the
cutter body. The major objection have been the initial cost and the difficulty in keep-
ing the cutting edges in the same plane. Many users have complained that index-
able-insert milling cutter do not produce acceptable surface finishes and are suitable
only for rough milling. However, continued research and development has produced
designs in construction and geometry that give very good surface finishes when used
Design of Cutting Tools 331

correctly. The initial cost is easily offset by the elimination of cutter sharpening and
maintenance.
The indexable-insert face-milling cutter consists essentially of a cutter body in which
are mounted a number of cutting inserts held in position by wedges or clamps are
mounted. The majority of manufactures feature wedge-behind-the-insert design. In
other words, hardened wedges are located behind the cutting inserts to hold them in
position. In case of an accident or collision, the wedge serves to absorb shock and
protect the cutter body. By the same token, many manufacturers use a replaceable
anvil or rest button to support the insert and to absorb shocks resulting from an acci-
dent. These features should be considered when purchasing a cutter, as it is more
economical to replace component parts than to replace an expensive cutter.
Figure 4.127 shows an example of a well-designed throwaway insert milling cutter.
The anvils of this particular cutter are welded to the cutter body, which has the effect
of marking it a solid, one-piece milling
cutter. The cutting inserts are completely
embraced by the hardened-steel anvil
pocket and are backed by a hardened-
steel wedge.
Many indexable-insert milling cutters
have a provision for adjusting the insert
in an axial direction. This feature enables
the operator to hold face runout to a mini-
mum, and the result is better finishes and
improved cutter performance. Using a
dial indicator it is possible to position the
cutting edges within ± 0.0025 mm in an
axial direction. This comes very close to Fig. 4.127 Component parts of a throw-
the same accuracy as solid ground cut- away-insert milling cutter
ters. Figure 4.128 shows a face-milling (Valenite Metal, Division of
cutter with this feature. The hardened- Valeron Corporation)
steel anvil provides positive location of
the insert yet can be moved axially to
remove runout. An added feature of this
cutter is that each anvil can provide axial
adjustment within itself if adjustable anvils
in each station of the cutter are specified.
A screw adjustment is provided in the
adjustable anvil to make one adjustment
possible without movement of the wedge
lock. Double-wedge design enables the
operator to index the insert by loosening
only one wedge.
A chip pocket is usually provided ahead
of the cutting, in the form of a concave
surface cut directly into the cutter body Fig. 4.128 Throwaway-insert milling
or an open space provided by seating cutter with anvils that can be
the wedge well below the periphery of adjusted in an axial direction
the cutter body. The chip pocket helps to (Lovejoy Tool Company, Inc.)
332 Tool Design

provide unobstructed clip clearance in front of the blade in order to allow chips to form
freely without back pressure. This is especially important when deep cuts and heavy
feeds are necessary.
Several different geometries involving combination of axial rake, radial rake, and lead
angles are available. Six basic rake geometries are available, depending upon the
standardisation of various companies. In other words, six geometries are available
but not all from single company.
The more popular rake geometries available are the double-negative, the double-pos-
itive, the positive-radial, negative-axial, negative-radial, and positive-axial all available
in varying degrees to provide chip control for different materials and setups.
Lead angles are available from 1 to 60°. In general, positive rake is recommended for
aluminium and nonferrous alloys, although there are exceptions. Combination posi-
tive-negative rakes have proved very successful for stainless steels. Double-negative
rakes offer greatest economy because of the added indexes per blade.
No hard-and-fast rules can be given for cutter and geometry selection. When in doubt,
the cutter manufacturer should be consulted before specifying a cutter. This past
experience enables him to set up guidelines that will give optimum results under
certain conditions.
As a general rule indexable-insert milling cutters are used for general-purpose rough
and semi finish milling applications where finish requirements are not critical. When
finish becomes important, many manufacturers offer an individual finishing or wiper
blade to smooth out the feed marks made by the milling-cutter teeth. This may take
the form of separate blades located inside the periphery of the regular cutting teeth,
or the shape of one of the cutting inserts may be altered to achieve the same pur-
pose. Cutters with provisions for removing the insert nest often utilise a wiper nest for
finish machining. One or more of the standard insert nests is replaced with a wiper
nest which automatically locates the wiper insert in the proper relation to the other
inserts.
Although each manufacturer uses his own special geometry to establish the relation-
ship of the wiper insert to the roughing inserts, finishing and wiper blades are based
on the principle of the broad-nose tool used for finish machining on the planer. The
cutting edge is parallel to the feed movement of the machine and is wide enough
to lap over several tool marks made by the roughing tools, therefore producing a
smoother finish (see Fig. 4.129).
When applied to a milling cutter, the wiper blade is set from 75 to 100 mm. higher than
the roughing teeth and has a flat section wide enough to overlap the tool feed marks
made in one revolution of the cutter. Thus, if the clip load per tooth is 0.25 mm and
there are 16 inserts in the cutter, the flat section on the wiper blade would have to be
over 4 mm in order to overlap the feed marks made by the roughing inserts with each
revolution.
Milling differs from planning in that the cutting edge is continually advancing into the
workpiece because of the continuous feed movement. Wiper blades are sometimes
designed with a slight lead angle (2 to 3°) to thin the clip gradually and blend the sur-
face produced by the roughing inserts with that produced by the wiping insert.
Design of Cutting Tools 333

Feed

Tool
0.05 TO 0.025

Workpiece

Feed
per
stroke Cutting edge parallel
to feed movement

Fig. 4.129 The principle of the broad-nose tool for finishing cuts on the planer or shaper

4.9 DRILLS AND DRILLING


The outward appearance of the standard twist drill has not changed during the past
50 years. Basically, it consists of a shank for mounting into the drilling machine, heli-
cal flutes that form rake angles and permit the escape of chips, and two cutting edges
separated by a chisel edge. The major improvement in recent times has been in the
accuracy during manufacturing and in the drill-point geometry, along with a shorter
and stiffer drill. These improvements have increased the performance of twist drills, but
users have been slow to adopt anything other than the old standard twist drill. Many
users believe that “drill are drills” as long as they are made of high-speed steel.
Drills are generally not considered a precision tool. One tool engineer expressed this
well when he said, “A drilling operation is usually done to allow air to pass through the
workpiece.” When size must be held to close limits, a precision sizing operation, such
as reaming or boring, usually follows a drilling operation.
Perhaps the twist drill is taken for granted because it has been around do long. It is
generally accepted that a twist drill will remove more metal per unit cost than any
other rotating cutting tool. With this in mind, the tool engineer should also consider
the improvement in performance through recent innovations in drill design. Drilling
may never be a precision hole-producing operation, but in many cases performance
can be improved to the point where reaming can be eliminated on workpieces that do
not require the ultimate in accuracy. The tool designer should be aware of the latest
improvements in drill design and geometry and should at least try them on an experi-
mental basis.

The drilling process Although the basic twist drill looks simple, the formation of
the chip it produces is quite complex because the chip is formed by two different tool
geometries of the outer cutting lips and the chisel edge. The cutting process along
the lips is essentially the same as that of single-point cutting tool. However, the cut-
ting process at the chisel edge is quite different. The chisel edge may be considered
as having an extremely negative value which tends to crowd the chip rather than cut
through it. Metal deformation in the area of the chisel edge is quite complex, and the
metal tends to be squeezed or extruded outward until it reaches a point where it is
picked up by outer cutting lips. The cutting action in this area is largely responsible for
334 Tool Design

the high axial-thrust forces required by large twist drills and explains the necessity of
pilot holes when using larger drills in smaller drilling machines. Figure 4.130 shows
the formation of a chip in the area of the chisel edge.

Fig. 4.130 Formation of a chip in the area of the chisel edge (National Twist Drill and Tool
Company)

The chisel edge is therefore one of the major problem areas in the geometry of twist
drills. Besides causing high axial-thrust forces, it may be responsible for locational
inaccuracies. Should one side of the chisel edge be slightly higher than the other,
the drill will tend to walk when it comes into contact with the workpiece. In the past,
this has generally been prevented by using guide bushings or by centre punching the
desired hole location. The development of numerically controlled drilling machines
has eliminated guide bushings and centre-punch marks, and it is therefore neces-
sary to use a drill point that will centre itself in the correct location without walking.
Drills used on numerically controlled drilling machines are discussed extensively in
Chap. 12.
The helical flute of a drill is another problem area not generally recognised. The major
purpose of the helical flute is to form the rake angle and provide a passage for chip
removal. Construction of this type results in a thin centre section connecting two
larger helical bodies. In a sense, the fluted section of a drill may be considered as a
torsion spring that is winding and unwinding. For example, a standard drill in the 3
to 5 mm range can be wound through an angle of 60 to 90° before it will break. This
winding action reduces the rigidity necessary at the cutting edge and may be directly
responsible for chattering in a drilling operation.
The tendency for a drill to wind is lessened by increasing its rigidity, either by decreas-
ing the flute length or increasing the web thickness. Decreasing the flute length is
probably more desirable, as the majority of drilling operations do not require the entire
flute length provided on a standard drill. When long-fluted drills are required, the web
thickness may be increased. This decreases the flute space, however, and the drill
may have to be withdrawn frequently to clean chips.
Drilled-hole size has already been discussed in part, and it is generally accepted that
the standard two-flute twist drill cannot be considered as a precision hole-producing
tool. It is commonly known that a drill will produce an oversize hole even when the
geometry is correct and it is properly used. The amount of oversize is fairly consistent,
and many toolmakers make use of this consistency.
Design of Cutting Tools 335

The amount of oversize depends upon the size of the drill. A knowledge of the
amount of oversize for a specific drill diameter is often important. The Metal Cutting
Tool Institute conducted a series of tests to determine the amount of oversize that a
drill may be normally expected to cut under normal shop conditions. The results are
shown in Fig. 4.131. This information is based on properly sharpened drills. Drills with
unequal cutting lips, improperly thinned web, and unequal lip angles may produce
exaggerated bell-mouthed and oversize holes.

Fig. 4.131 The amount of oversize a drill will cut under normal shop conditions. (From
MCTI, “Metal Cutting Tool Handbook”)

Nomenclature of twist-drill elements The Metal Cutting Tool Institute lists the fol-
lowing important twist-drill elements, as shown in Fig. 4.132. A complete listing may
be found in the “Metal Cutting Tool Handbook.”

Tang Point angle

Taper shank Clearance


Drill diameter diameter
Tang Body dia. clearance
drive
Chisel edge
Lip relief angle
Flutes angle
Helix
Neck angle Margin
Shank Straight
Axis
diameter shank

Shank Flute length Lip Land


length Body Web
Overall Chisel edge
length

Fig. 4.132 Terms applying to twist drills (From MCTI, “Metal Cutting Tool Handbook”)
336 Tool Design

Axis The imaginary straight line which forms the longitudinal centreline of the drill.
Back taper A slight decrease in diameter from front to back in the body of a drill.
Body The portion of the drill extending from the shank or neck to the outer corners
of the cutting lips.
Chip breaker A nick or groove designed to reduce the size of chips. It may be a
step or groove in the cutting lip or in the leading face of the land at or adjacent to the
cutting lips.
Chisel edge The edge at the end of the web that connects the cutting lips.
Chisel-edge angle The angle included between the chisel edge and the cutting lip,
as viewed from the end of the drill.
Clearance The space provided to eliminate undesirable contact between the drill
and the workpiece.
Clearance diameter The diameter over the cutaway portion of the drill lands.
Cutting edges See Lips, the preferred term.
Drill diameter The diameter over the margins of the drill measured at the point.
Flutes Helical or straight grooves cut or formed in the body of the drill to provide cut-
ting lips, to permit chip removal, and to allow cutting fluid to reach the cutting lips.
Heel The trailing edge of the land.
Helix angle The angle made by the leading edge of the land with a plane containing
the axis of the drill.
Land The peripheral portion of the body between adjacent flutes.
Lips The cutting edges of a two-flute drill extending from the chisel edge to the
periphery.
Lip relief The axial relief on the drill point.
Lip relief angle The axial relief angle at the outer corner of the lip. It is measured by
projection into a plane tangent to the periphery at the outer corner of the lip.
Margin The cylindrical portion of the land which is not cut away to provide
clearance.
Neck The section of reduced diameter between the body and the shank.
Point The cutting end of a drill, made up of the ends of the lands and the web. In
form it resembles a cone, but it departs from a true cone to furnish clearance behind
the cutting lips.
Point angle The angle included between the cutting lips projected upon a plane
parallel to the drill axis and parallel to the two cutting lips.
Shank The part of the drill by which it is held and driven.
Tang The flattened end of a taper shank intended to fit into a driving slot in a
socket.
Design of Cutting Tools 337

Web The central portion of the body that joins the lands. The extreme end of the
web forms the chisel edge on a two-flute drill.
Web thickness The thickness of the web at the point, unless another location is
indicated.

4.10 TYPES OF DRILLS


According to the formal definition often quoted, “Twist drills are end-cutting tools with
one or more cutting edges having helical or straight flutes or grooves adjacent thereto
for the admission of a coolant and the ejection of cutting or chips.” Standard twist drills
are available in a variety of designs. A standard drill is one that is commonly carried in
stock by distributors. On the other hand, drills that require special forms, dimensions,
and tolerances are referred to as special-purpose, in that they are not carried in stock
and must be made to specifications. It is often hard to distinguish between standard
and special-purpose drills because a drill that will be special-purpose to one manufac-
turer may be standard to another. A few companies specialise in special-purpose drills
and carry them as standard stock items. The tool or manufacturing engineer should
be familiar with the various sources of drills because it is often possible to purchase
special drills at less cost from specialty companies. The various types of standard and
special-purpose drills are discussed in the following paragraphs:

Straight-shank drills These drills have straight shanks the same diameter as the
drill bodies. The common variables in manufacture are drill length, helix angle, and
web thickness. They range in size from approximately 0.15 to 50 mm, depending
upon their construction. The most common is the jobber’s-length straight-shank drill.
It is a general-purpose drill designed to perform satisfactorily under as many normal
conditions as possible (see Fig. 4.133a). When conditions call for the use of short
drills, a screw-machine length may be used. These drills were primarily developed for
use in screw machines but may be used in other drilling operations when maximum
rigidity is required, as in portable drilling and sheet-metal drilling. They also have
found wide use in drilling tough, hard materials when short rigid tools are a basic
requirement (see Fig. 133b). Centre drills are similar in appearance to the smaller
sizes of screw-machine-length drills except they are shorter. They are very short
overall and are designed for centrtreing work. Standard diameters of centre drills are
from 1.5 to 8 mm.
Closely related to the centre drill is the spotting and centreing drill, shown in Fig.
1.133c. The main difference is that the standard diameters of spotting and centreing
drills range from 10 to 25 mm and no body clearance is provided behind the drill mar-
gins. This allows chucking close to the drill point for accurate drill starting or centreing.
Spotting or centreing drills are designed to produce accurate and true centres in all
types of screw-machine work. This is possible because they provide alignment and
true centres for starting follow-up drills. A constant web is utilised that requires no
thinning when repointing.
Taper-length general-purpose straight-shank drills are patterned after the standard
taper-shank drill except that they are made with a straight shank, not flattened, of the
same diameter as the fluted section. It is a general-purpose drill somewhat longer
than jobber’s-length straight-shank drills (see Fig. 4.133d).
338 Tool Design

Certain styles of standard straight-shank drills have a tang drive (Fig. 4.133e) which
permits the use of a split-sleeve drill driver under severe drilling operations. They can
also be used in three-jawed or conventional drill holders as well. Straight-shank drills
with a tang drive are often referred to as automotive-series straight-shank drills.

Taper-shank drills These drills have a standard taper shank and are designed for
general-purpose drilling. The size ranges from 3 to 90 mm. The standard taper is the
Morse system. Heavy-duty taper-shank drills are available when the severity of drilling
conditions demands stronger tools. Figure 4.133f shows a standard taper-shank drill.

Extra-length drills These drills are designed for drilling in inaccessible places or for
drilling holes of extreme depths. Aircraft extension drills with spring-tempered shanks
consist of a standard drill with either a 150 or 300 mm extended shank. They are used
principally for drilling holes in inaccessible places which cannot be reached with regu-
lar-length drills. Many such holes are encountered in aircraft-assembly work on wings,
spars, firewalls, cowling, etc., where holes are made with portable air or electric drills.
The spring-tempered shank permits drilling not easily reached in direct alignment.
Extra-length drills for deep-hole drilling are drills with extra flute length. They are usu-
ally of heavy web construction to increase the drill rigidity, with a split-point type of
web thinning. The helix angle of the flute is generally higher than normal to increase
the efficiency of chip removal. Figure 4.133g and h show typical examples of extra-
length drills.

High-helix drills This design was developed for drilling deep holes in materials of
low-tensile strength such as aluminium, magnesium, copper, die-casting materials,
wood, plastic, slate, and marble. Heavy feeds may be used on materials of this type,
and the result is a chip-disposal problem. High-helix drills utilise wide polished flutes
and high helix angles to help clear chips (see Fig. 4.133i).

Low-helix drills Drills of this type have been designed primarily for the production
drilling of brass but are also used for shallow drilling some aluminium and magnesium
alloys. The rake angle of the cutting edge is reduced with the reduction of the helix
angle, and thus the tendency of the drill to grab or “hog in” is reduced during heavy
feeding. The extreme of this offset is in straight-fluted drills, which will not run ahead
or grab because of their zero rake angle. They are also available for use in thin sheet
material for the same reason (see Fig. 4.133j).

Three- and four-flute core drills These drills are used for enlarging holes that
have been previously drilled, cored, or punched. Because of their use in cored holes
in castings, they are commonly called core drills. They are pointed so that they will
enlarge holes of approximately 60 per cent of the tool diameter. The advantages of
tools of this type are increased productivity, better finish, and greater accuracy of hole
size and location. Increased productivity is gained through increased feeds because
of the extra cutting edges. Core drills also have a heavy centre section that will stand
the additional strain imposed by the greater number of cutting lips. The finish is better
because the additional lands tend to support the drill more evenly around the circum-
ference of the hole. The result is less tendency to wobble, to score up the walls of the
drilled hole, or to cut oversize.
It has been said that the core drill is halfway between a drill and a reamer. It has
many of the characteristics of both. The finished hole is of better quality than a drilled
Design of Cutting Tools 339

hole but inferior to that of a reamer. Core drills have replaced reamers when a large
amount of material must be removed. Figure 4.133k shows a typical core drill.

Multiple-diameter drills Drills of this type produce a hole of two or more diameters
with a single drill (see Fig. 4.133l). The most common of the multiple-diameter drills
is the step drill, which has two or more diameters produced by grinding various suc-
cessive steps on the lands of the drill. These steps are usually separated by square or
angular cutting edges, as determined by the application. The type is extensively used
because it can be made by grinding down and stepping an ordinary drill. Some web
thinning is usually required when this procedure is followed.

(a) (b) (c) (d) (e) (f) (g)

Mohawak
Standard
Subland drills

Subland drill-Chamfer
for Regular Tops-Flute-
les Taps-Pipe Tapes-Coil
insert Taps

Subland drill-Couater-
bete- Socket Head-
Cap Screws

Mohawk Standerd Subland Drill-Drill-for


Sublands are stocked Top Drill and Body
in three basic types-in Clearance
three step lengths and
three shank styles all
practical sizes and Mohawk Tools. INC.
Montpsecise OHIO
(h) (i) (j) (k) combinations (l)

Fig. 4.133 Types of drills: (a) Jobber’s-length drill, (b) Screw-machine-length drill, (c) Spot-
ting and centreing drill, (d) Taper-length straight-shank drill, (e) Straight-shank
drill with tang drive, (f) Taper-shank drill, (g) Straight-shank extra-length drill,
(h) Aircraft extension drill, (i) High-helix drill, (j) Low-helix drill, (k) Four-flute
core drill, (l) subland drill. [(a) to (k): The DoAll Company; (l): Mohawk Tools,
Inc.]
340 Tool Design

The disadvantage of step drills is in the resharpening because the cutting shoulder
as well as the point must be reground in order to maintain the correct length for each
diameter. Regrinding the shoulders almost always results in nicking the next smaller
diameter. After repeated resharpenings, a point is reached where the drill can no
longer be resharpended; it can be salvaged only by a complete reestablishment of a
new, smaller diameter.
The subland drill, sometimes called the multicut or dual-cut drill, overcomes the
difficulty in resharpening. (This type performs the same functions are the step drill
described above, though its somewhat different. The step drill has its steps on differ-
ent diameters on the same land, while the subland drill has two distinct lands running
substantially the entire length of the flutes.) The advantage is that the two diameters
can be maintained constant throughout the life of the drill because in resharpening
the cutting edges of the large diameter it is not necessary to touch the margins of the
small diameter. It is somewhat more expensive to manufacture than the step drill, but
the convenience and economy in resharpening often outweigh the difference in first
cost. Subland drills, like step drills, may be made in more than two diameters and are
not necessarily restricted to two-fluted tools.

Drill-point geometry The geometry of the twist drill is determined by the manufac-
turer and cannot be altered to any great extent by the user. The geometry of the drill
point, however, can be altered to suit specific needs. Alterations may be necessary
for proper chip control, to increase cutting efficiency, or to increase the accuracy of
hole location.
Drill-point geometry is made up of the helix angle, the point angle, and the lip clear-
ance angle. An alteration of any one of these elements will effect the performance of
a drill.
The helix angle of the flutes determines the rake angle of the drill. This is a general
statement because there are other factors that determine the true-rake, or effective-
rake, angle. The helix angle also serves to reduce axial-thrust and torque forces.
For general-purpose high-speed-steel drills the helix angle ranges from 22 to 33°,
depending upon the drill diameter.
A closer analysis of the rake angle reveals that it is not constant along the entire cut-
ting edge. Normal to the cutting edge it is positive near the outer corners of the cutting
edge and gradually decreases toward the web. Near the web the normal rake angle
becomes negative due to the rapid change of shape near the bottom of the flute. This
means that a portion of the chip would be formed under positive-rake conditions and
another part near the web would be formed under negative-rake conditions, assuming
that chip flow was normal to the cutting edge. However, it is now believed that chip
flow is not normal to the cutting edge, as has been assumed, because the cutting
edges near the web have a high inclination to the direction of motion due to the point
angle and produce a shearing cut. This means that the chip is deflected through a
lesser angle than it would be if it flowed normally to the cutting edge. The effective
rake then becomes positive along the entire length of the cutting edges. The chip flow
is somewhat similar to that obtained with single-point tools having combined positive
and negative side and back rake to form a positive true rake.
The preceding discussion would indicate that the effective rake could be changed
by changing the included drill-point angle. This variable is easily changed with any
Design of Cutting Tools 341

modern drill-point grinder. An increase in the included point angle would increase
the effective-rake angle, as shown in Fig. 4.134. A drill having a comparatively small
included point angle has a small effective rake angle. The cutting efficiency of the
drill can therefore be increased by increasing the included point angle because the
effective rake angle approaches the helix angle of the drill. However, too great an
increase in the included point angle is not advisable, because the decrease included
point angle will require abnormal axial thrust for proper feeding. It also has a greater
tendency to walk when starting a hole without guide bushings.

Effective
rake angle

Effective
rake angle

Fig. 4.134 Changing the effective rake angle of drills by changing the included drill-point
angle

The standard included drill-point angle is 118°. It is used for drilling a wide variety of
materials, including mild steels, cast irons, and many alloy steels. In production drill-
ing, better results are often obtained by increasing the included point angle from 135
to 140° because of the increased effective-rake angle. This is especially true when
drilling deep holes or hard and tough materials. Controlled feeds and speeds should
be used, along with guide bushings. The flat angle point has a tendency to walk as it
contacts the workpiece.
The long angle point (60 to 90°) is commonly used on soft plastics, wood, and other
soft nonferrous materials with a low resistance to penetration. Drills entering this type
of material sometimes have a tendency to act like a corkscrew and suck into the
workpiece. The long point angle reduces the effective-rake angle and retards this ten-
dency to some extent. Low-helix drills are often used in combination with long angle
points to further reduce the effective-rake angle. Table 4.21 gives the recommended
point angles for common materials.
The corners located on the outer extremities of the cutting edge are subjected to
more abrasion and wear than any other portion of a drill. Corners on drills with large
included point angles have small corner angles and therefore are more susceptible to
abrasion than drills with small included point angles. The abrasion area may be further
increased by grinding a secondary angle at the corners of the drill, as shown in Fig.
4.135e. This type of point is especially well suited when abrasive materials are drilled,
such as medium and hard cast irons. The double angle acts as a chip breaker when
drilling steels because of additional chip distortion it causes.
A large amount of thrust is required to force the chisel edge of a drill into the work-
piece. Any modification that will reduce the chisel edge will improve the drill’s effi-
ciency. “Splitting the point” is one common method of reducing the chisel-edge area.
Table 4.21 Recommended drill-point geometry for various materials
Material Hardness, BHN Point angle, LP relief Chisel-edge Helix angle, Point grind*
(Rockwell values in angle, deg angle, deg deg
parentheses)
342 Tool Design

Free-machining plain carbon steels, 100-125 118 12-15 125-135 24-32 S


plain carbon steels, free-machining 225-325 118 10-12 125-135 24-32 S
alloy steels, alloy steels, nitriding 325-425 118-135 8-10 125-135 24-32 C
steels, armour plate, tool steels, RC45-52 118-135 7-9 125-135 24-32 C
cast steels
Ultrahigh-strength steels 175-425
RC45-52 118 8-10 125-135 24-32 C
110-225 118 8-12 125-135 24-32 S
Grey, ductile and malleable irons
225-400 118 10-12 125-135 24-32 S
135-200 118 10-12 125-135 24-32 S
Ferritic, austenitic, martensitic, and
precipitation-hardening stainless 200-325 118 7-10 125-135 24-32 C
steels 325-425
RC48-52 118-135 7-10 120-130 24-32 C
Titanium alloys 110-440 118 7-10 125-135 24-32 C
High-temperature alloys 140-400 118-135 9-12 125-135 24-32 C
Tungsten alloys 180-320
Molybdenum, columbium, and tantalum alloys 170-290 118 7-10 125-135 24-32 S

Contd...
Contd...

Material Hardness, BHN Point angle, LP relief Chisel-edge Helix angle, Point grind*
(Rockwell values in angle, deg angle, deg deg
parentheses)

Nickel alloys 80-360 118 8-12 125-135 24-32 C


Nitinol alloys 210-360
RC48-52 118 7-10 125-135 24-32 C
Aluminium alloys 30-150 90-140 12-15 125-135 24-48 S
500 kg
Magnesium alloys 40-90 70-118 12-15 120-135 10-30 S
500 kg
Copper alloys RE20-100 118 12-15 125-135 10-30 S
Zinc alloys 80-100 118 12-15 120-135 24-32 S
NOTE: Use stub-length drills whenever possible on high-strength materials.
*S = standard; C = crankshaft.
SOURCE: “Machining Data Handbook,” Metcut Research Associates, Inc.
Design of Cutting Tools 343
344 Tool Design

The splitting operation actually creates two new secondary cutting edges which
extend almost to the centre of the drill, eliminating most of the chisel edge and its
negative rake effect. The split-point drill is often referred to as the crankshaft point
because it was originally developed for use on drills designed for drilling deep holes in
automotive crankshafts. The extra-long drills required a heavy web to provide rigidity
and a means of reducing the resulting wide chisel edge. In addition to a reduction in
thrust, the split point minimises skidding or walking of the drill point when starting a
hole. Figure 4.135d shows the split or crankshaft point.
The advent of numerically controlled drilling machines placed new demands upon
drill design and especially upon drill-point geometry. A numerically controlled drilling
machine that will position within 25 mm or less requires drills capable of producing
similar accuracy. Since a guide bushing is not used with numerically controlled drill-
ing machines, it is imperative that the drill does not walk as it contacts the workpiece.
Several new drill-point geometries have been developed to meet this need, such as
the spiral-point drill. Drills used with numerically controlled machines are discussed
in detail in Chap. 12.

120°-135°
135°

8°-12°
12°e
118°
90°
(a) (b)

115°-125° 125°

6°-9° 9°
Secondary
135° angle
128°

32° (e)
(c)
(d)

Fig. 4.135 Common types of drill points: (a) Standard (b) Long-angle (c) Flat (d) Split or
crankshaft (e) Double-angle

Chip control in drilling operations Materials which produce a discontinuous chip


generally present no problem with regard to chip control. However, when drilling mate-
rials that produce a continuous chip, it is often desirable and sometimes necessary to
break the chip into short lengths for easy removal. Too many man-hours are wasted
while the operator stops the drilling machine to untangle long chips wound around
the drill. A mass entangled chips also endangers the operator, as well as knocking a
flexible coolant nozzle out of position. When this occurs, some means of breaking the
chip should be found.
Design of Cutting Tools 345

Breaking the chip may be a simple matter of increasing the feed rate, using a double
point angle, or grinding a small flat along the cutting edges. This of course reduces
the rake angle, and the same effect can be obtained by using drills with less helix
angle, provided the depth of the hole is not too deep. If the chip still does not break, it
may be necessary to grind a groove behind the cutting edge as shown in Fig. 4.136b.
However, this requires special skill and is time consuming. Perhaps, a better approach
would be to use a drill with a built-in chip-breaker, as shown in Fig. 4.136a. Here a
small radius is built into the flute shape and extends the entire length of the flute to
curl the chip tightly. The resultant chip formation breaks up the material into small seg-
ments. Drill points are reground in the conventional manner. Chip-breaker drills are
available from drill specialty companies as standard items, and the small increased
cost will be offset by increased productivity.

Curled chip Radius built into


flute to force
chip curl
Ground-in
groove

(a)
(b)

Fig. 4.136 Methods of breaking chips during drilling operations: (a) Drill with built-in chip
breaker. The radius extends the entire length of the flute (b) Chip breaker ground
into cutting edge of conventional drill

Troubleshooting drilling problems The following common types of drill failure,


along with the cause and remedy, may be used as a guide for average shop condi-
tions. Specific problems of an unusual nature may require the advice of a representa-
tive from the drill manufacturer.
1. The outer corners are broken or appear to be “wiped off.” This is an indication
that the cutting speed is too high, especially if the drill blues at the outer cor-
ners. Insufficient or improper lubricant is also a possible cause, along with hard
spots, scale or sand inclusions in the material being drilled.
2. Oversize holes may be caused by cutting lips with unequal angles or off-
centre drill points resulting from improper web thinning or drill reconditioning.
Eccentricities in the drill itself or the machine spindle may also cause oversize
holes.
3. Cutting lips chipped along the entire length may be the result of excessive feed
or too much lip relief. Excessive lip relief results in lack of support behind the
cutting edge.
4. A drill spilt up the web indicates that abnormal thrust was required to make the
drill penetrate the workpiece. Causes may be excessive feed, inadequate lip
relief, drill point thinned too much, or light constructed drill. A heavy-duty drill
with a thicker web may be required. A split point may also solve the problem.
346 Tool Design

5. Poor chip removal will cause loading and galling and usually occurs when drill-
ing soft materials like aluminium, magnesium, copper, and thermoplastics,
materials which produce a large volume of chips that pack into the flutes. A
fast-helix drill will usually help, as the low angle of incline of the flutes is espe-
cially suited for their removal. Inadequate application of cutting fluid may also
be the cause. In cases where the cutting fluid cannot get into the flutes, a drill
with a specially treated surface may help. See surface treatments for metal-
cutting tools.
6. Improper drill reconditioning is one of the major factors contributing to poor
drill performance. A drill should be given the same consideration during the
resharpening operation as a milling cutter. The drill point should be resharp-
ened on a precision machine designed for that purpose. Few toolmakers or
machinists can recondition a drill to its original condition as it came from the
manufacturer.

Power requirement for drilling It is sometimes desirable to have a general idea of


the amount of thrust (feeding pressure) required to feed a drill and the power required
to drive a drill. A knowledge of feeding pressures required to feed a given size drill
aids the cool designer in determining the required strength of drill jigs. A knowledge
of power requirements for drilling is of value when planning drilling operations or pur-
chasing new equipment. Basic formula are available for determining feeding pressure
and power requirements, but their use in time-consuming. The Metal Cutting Tool
Institute has published tables which makes the task much easier (see Tables 4.22
and 4.23). Table 4.22 is used for determining the thrust in KN required for drilling AISI
1112 steel with a sharp drill with conventional point and web dimensions. No allow-
ance was made for drill-point wear, which may increase thrust by 100 per cent or
more. Approximate conversion factors are given for thrust to drill other materials.
Table 4.23 gives the power required for drilling AISI 1112 steel at 100 rpm. No allow-
ance was made for drill-point wear, for losses in power transmission to the spindle of
the drill press, or for the small amount of power required to feed the drill into the work.
The Metal Cutting Tool Institute therefore recommends that a minimum of twice the
values given in Table 4.23 be used in making equipment estimates.

4.11 REAMERS
Drills are not a precision hole producing tool. When size must be held to close limits or
the surface finish is important, a reaming operation usually follows a drilling operation.
Reaming is not as accurate as single-point boring because it tends to follow the path
of the drilled hole, but it does improve the accuracy of a round hole enough to allow
interchangeability. Reamers hold size consistently and can be applied to the metal
machining process with a minimum amount of skill.

The reaming process Reaming is a finishing operation and is not intended to remove
a large amount of material. However, enough stock should be removed to allow the
cutting edge to “get under the chip” or the reamer will burnish the work rather than cut
it, especially when the reamer starts to dull. In general terms, the amount of material
to be removed for machine reaming is 0.13 to 0.25 mm up to a 6 mm hole, 2.5 to 0.4
mm up to a 13 mm hole, and 0.4 to 0.8 mm up to a 40 mm hole. Hand-reaming stock
allowances range from 25 to 130 mm because of the difficulty in forcing the reamer
through the workpiece material.
Table 4.22 Thrust in kN for drilling AISI 1112 steel
Drill size mm Feed, mm/rev
0.025 0.051 0.10 0.15 0.2 0.25 0.33 0.41 0.51 0.64 0.76

3 0.10 0.19 0.33 0.48 ... ... ... ... ... ... ...

6 0.20 0.36 0.64 0.96 0.18 1.45 ... ... ... ... ...
10 ... 0.53 1.20 1.36 1.73 2.11 ... ... ... ... ...
12.5 ... 0.67 1.20 1.71 2.18 67 3.29 3.96 3.96 ... ...
16 ... ... 1.42 2.00 2.60 3.11 3.91 4.67 5.78 ... ...
19 ... ... 1.67 2.36 2.93 3.64 4.59 5.34 6.67 ... ...
22 ... ... 1.81 2.71 3.38 4.22 5.23 6.25 6.25 ... ...
25 ... ... ... 3.11 3.95 4.89 6.00 7.34 8.90 ... ...
32 ... ... ... 3.91 4.89 6.23 7.78 9.34 11.39 13.79 ...
38 ... ... ... 4.89 6.22 7.78 9.79 11.57 14.23 17.34 ...
45 ... ... ... 6.00 7.28 9.56 12.01 14.46 17.79 21.12 ...
51 ... ... ... 7.34 9.34 11.56 14.45 17.79 21.35 25.80 30.25
57 ... ... ... 8.89 11.34 14.23 17.79 21.35 25.80 31.13 36.46
64 ... ... ... 10.68 13.79 16.90 20.91 25.80 31.13 37.36 43.59
70 ... ... ... 12.45 16.08 20.01 24.91 30.25 36.48 44.48 52.27
76 ... ... ... 14.68 19.13 24.02 29.36 42.70 35.59 51.15 60.05
Contd...
Design of Cutting Tools 347
Approximate conversion factors for thrust to drill other materials
Material Thrust
Regular points Split points
348 Tool Design

MST 6A1– 4VA titanium alloy BR 340 2.0 0.8


17 – 7 PH stainless steel BR 400 2.3 1.0
4340 Heat-treated steel 240,000 to 260,000 psi 3.5* 1.6
AISI 1020 1.4
AISI 1035 1.3
1.00% C tool steel 1.7
AISI 3150 1.4
Malleable iron 0.6
Gray cast iron 0.6
Example:
Find the thrust developed in drilling malleable iron with a 16 mm drill at a feed of 0.33 mm/rev
From thrust listing, 16 mm drill at 0.33 mm/rev feed = 3.19 kN
Conversion factor for a malleable iron = 0.6
Converting to thrust for malleable iron = 3.91 × 0.6 = 2.35 kN
* Regular points with special thinning are used for drilling this material.
SOURCE: The Metal Cutting Tool Institute.
Table 4.23 Power required for drilling, in watts AISI 1112 steel at 100 RPM
Drill size, Feed, mm/rev
mm 0.025 0.051 0.1 0.15 0.2 0.25 0.33 0.41 0.51 0.64 0.76

3 22 3 5 7
6 75 10 17 22 28 31
19 12 20 36 507 597 701
12.5 20 31 60 86 101 122 15 187 21
19 45 75 134 187 239 284 336 388 463
25 75 127 231 313 388 477 567 671 821 970
32 112 194 336 477 582 701 858 1007 1194 1417
38 164 269 463 657 821 970 1194 1417 1641 1940
45 224 373 612 895 1119 1268 1641 1865 2238 2611
51 276 463 821 1119 1417 1641 2014 2387 2835 3357 3879
64 ... ... ... ... 2089 2462 2984 3581 3954 5147 5819
76 ... ... ... ... ... ... ... 5147 5968 6789 8206

Approximate conversion factors for power required to drill other materials


Material Factor

AISI 1020 1.6


AISI 10.35 1.3
1.00% C tool steel 1.7
AISI 3150 1.6
Gray cast iron 0.5
Malleable iron 0.6
Contd...
Design of Cutting Tools 349
Approximate conversion factors for power required to drill other materials

Material Factor
Stainless steel:
350 Tool Design

AISI 416 free-machining martensitic 1.2


AISI 303 free- machining austenitic 1.6
AISI 304 austenitic 1.8
17 – 7 PH precipitation-hardened austenitic 2.0
4340 steel heat-treated to 1655-1792 MPa 2.3
Example:
Find the power required to drill a 12.5 mm hole in 416 stainless at 75 rpm with a feed of 0.15 mm/rev
Power at 100 rpm for 12.5 mm. drill at 0.15 mm/rev feed = 86 W
86
____
Power at 1 rpm = = 0.86 W
100
Material factor for 416 stainless = 1.2
Power at 75 rpm = 0.86 × 75 × 1.2 = 77.4 W
SOURCE: The Metal Cutting Tool Institute.
Design of Cutting Tools 351

The speed for reaming is generally considered to be two-thirds of the speed used for
drilling the same material. Reamers are more easily damaged than drills when sub-
jected to chatter. Since rotating cutting tools are less likely to chatter when the cutting
speed is reduced, reamers generally produce better results at lower cutting speeds.
Reaming speeds should always be reduced when chatter occurs. Finish reaming for
close tolerances and fine finish may require considerably slower speeds.
Reaming feeds are much higher than drilling feeds because there are more teeth in
the reamer. Feeds vary from 0.04 to 0.13 mm per flute per revolution, depending upon
the size of the reamer. Reamer feeds that are too low may cause burnishing, glazing,
or chatter. High feeds tend to reduce the accuracy of the hole and the quality of the
finish.
Practically all cutting by machine reamers is done by the chamfer which forms a trun-
cated cone on the end of the reamer. The chamfer angle generally used is 45°. Little
or no cutting is done on the longitudinal cutting edges behind the chamfer.
In contrast, cutting by a hand reamer is along the longitudinal cutting edge. The chip
produced is thinner and wider than that produced by the end-cutting machine reamer.
The hand reamer removes very little material, and its scraping action will produce a
very accurate hole when used by a skilled workman. Hand reamers were originally
used to finish and accurately size previously bored or machine-reamed holes for fitting
bushings or bearings. The precision honing machine has largely replaced the hand
reamer for this purpose, and it is now basically used for all special fitting applica-
tions. However, the accuracy and cutting action of a hand reamer has implications for
machine-reamer design.
The machine reamer in its simplest form has teeth chamfered on the end and cuts
only on the end. The longitudinal teeth are not relieved and form a circular land nearly
as wide as the flutes, as shown in Fig. 4.137. The diameter near the shank end is
slightly smaller than at the front to provide longitudinal relief. This type of machine
reamer is known as the rose chucking type and is suitable for heavy roughing cuts.
It is not accurate and will cut oversize. Rose reamers, therefore, any usually made
0.08 to 0.13 mm under nominal size. Rose reamers are used where a heavy cut is
necessary, as in the cored hole of a casting. A finish reaming operation follows a rose
reamer when extreme accuracy is necessary.

Circular
land

45°

(a) (b)

Fig. 4.137 (a) Basic construction of rose reamer (b) End view
352 Tool Design

Relieving the land of the rose reamer to provide a longitudinal cutting edge will
increase the accuracy and produce a reamer capable of finishing holes accurately
and smoothly. However, relieving the land also weakens the cutting edge, and the
amount of material to the removed must be reduced. The majority of the cutting is
still on the chamfered end, but the relieved land will size a hole by scraping away
roughness caused by the dulling of the peripheral corners of the chamfer. This type
of machine reamer, called a fluted chucking reamer, and will remove from 0.005 to
0.010 in of material where extreme accuracy is required.
An additional modification of the machine
reamer will improve the surface finish
and accuracy of the hole. A hand-reamer
grind is sometimes applied to obtain the
required finish or tolerance (see Fig.
4.138). In this case, a secondary cham-
fer of 1 to 10° is ground immediately back
of the 45° chamfer. Relief is provided on Fig. 4.138 Hand-reamer grind applied to a
this taper in a manner that will bring it to a machine reamer to obtain fine
sharp edge. The secondary chamfer thins finish and close tolerance
the chip and provides a scraping cut that
helps to size and smooth the walls of a hole in the same manner as a hand reamer.
This type of grind should remove less material than the standard fluted chucking
reamer for best results.

Nomenclature of reamer elements The Metal Cutting Tool Institute gives a com-
plete listing of reamer elements. The important elements are listed below and shown
in Fig. 4.139.
Back taper A slight decrease in diameter, from front to back, in the flute length of
reamers.
Blade A tooth or cutting element inserted in a reamer body. It may be adjustable
and/or replaceable.
Body The fluted full-diameter portion of a reamer, inclusive of the chamfer, starting
taper, and bevel. The principal supporting member for a set of reamer blades. Usually
including the shank.
Chamfer The angular cutting portion at the entering end of a reamer.
Chamfer angle The angle between the axis and the cutting the edge of the chamfer
measured in an axial plane at the cutting edge.
Clearance The space created by the relief behind the cutting edge or margin of a
reamer.
Cutting edge The leading edge of the land in the direction of rotation for cutting.
Flutes Longitudinal channels formed in the body of the reamer to provide cutting
edges, permit passage of chips, and allow cutting fluid to reach the cutting edges.
Angular flute A flute which forms a cutting face lying in a plane intersecting the
reamer axis at an angle. It is unlike a helical flute in that it forms a cutting face which
lies in a single plane.
Design of Cutting Tools 353

Overall length
Shank length

Tang
Flute length
Taper shank
Cutter
sweep
Chamfer
angle Chamfer
length
Helix angle
Straight shank Actual
size
Helical flutes
Shank length right-hand Body
helix shown
(a)

Starting
Neck taper Neck
Overall length
Shank length Flute length Pilot
Guide
Axis

Squared shank Cutter sweep Actual size


Cutter sweep
straight flutes
(b)

Starting Chamfer-relief
taper Chamfer length angle
Margin Relieved land Land
Cutting edge Relief angle Margin width
Land Flute Heel
width Actual Cutting Actual
size face size
Land

Core diameter Chamfer Radial


Bevel Chamfer angle
Helical flutes relief rake angle
left-hand helix shown straight flutes shown

(c) (d)

Fig. 4.139 Elements of metal-cutting reamers. (a) Chucking reamer Straight and taper
shank (b) Hand reamer Pilot and guide (c) Hand reamer; 0° radial rake angle
and right-hand rotation shown (d) Machine reamer; positive radial rake angle
and right-hand rotation shown (From MCTI, “Metal Cutting Tool Handbook”)

Helical flute A flute which is formed in a helical path around the axis of a reamer
(sometimes called spiral flute).
Heel The trailing edge of the land in the direction of rotation for cutting.
354 Tool Design

Helix angle The angle which a helical cutting edge at a given point makes with an
axial plane through the same point.
Land The section of the reamer between adjacent flutes.
Neck A section of reduced diameter connecting shank to body or connecting other
portions of the reamer.
Periphery The outside circumference of a reamer.
Pilot A cylindrical portion preceding the entering end of the reamer body to maintain
alignment.
Rake The angular relationship between the cutting face or a tangent to the cutting
face at a given point and a given reference plane or line.
Axial rake Applies to angular (not helical or spiral) cutting faces. It is the angle
between a plane containing the cutting face or tangent to the cutting face at a given
point and the reamer axis.
Helical rake Applies to helical and spiral cutting faces only (not angular). It is the
angle between a plane tangent to the cutting face at a given point on the cutting edge
and the reamer axis.
Hook A concave condition of a cutting face. The rake of a hooked cutting face must
be determined at a given point.
Negative rake Describe a cutting face in rotation whose cutting edge lags the sur-
face of the cutting face.
Positive rake Describe a cutting face in rotation whose cutting edge leads the sur-
face of the cutting face.
Radial rake angle The angle in a transverse plane between a straight cutting face
and a radial line passing through the cutting edge.
Shank The portion of the reamer by which it is held and driven.
Squared shank A cylindrical shank having a driving square on the back end.
Starting taper A slight relieved taper on the front end of a reamer.
Tang The flattened end of a taper shank which fits a slot in the socket.
Total indicator variation (TIV) The difference between the maximum and minimum
indicator readings obtained during a checking cycle.

4.12 REAMER CLASSIFICATION


The board general classification of reamers is into hand and machine reamers. Hand
reamers have been discussed to a limited extent and will not be mentioned again,
except in general terms, because of their limited use. The various common types of
machine reamers are noted in some detail because of their large-scale use. The tool
designer should be aware of the advantages and disadvantages of standard ream-
ers, since he will probably be more concerned with selection than with actual reamer
design, which for the average tool engineer is limited to modifying standard designs.
Design of Cutting Tools 355

Chucking reamers Chucking reamers are designed primarily for application in tur-
ret lathes, drill presses, screw machines, and automatic lathes. They are suited for
reaming the common materials ordinarily encountered. The standard chamfer angle
is 45°, although it may be modified to suit certain conditions. Chucking reamers are
readily available with straight flutes or right- or left-hand spiral flutes in either the
fluted chucking type or the rose chucking type. A spiral-fluted reamer generally has
less tendency to chatter and should always be used when cutting through an existing
hole in the side wall of the parent hole or when reaming holes containing slots or key-
ways. Right-hand spiral reamers cut more freely than straight-fluted or left-hand spiral
reamers, but they have a tendency to hog into soft materials such as certain types
of brass. Left-hand spiral reamers tend to push back against the feeding mechanism
of the machine too and are well suited for loose linkage. In other words, they keep
any slack or backlash in the feeding mechanism in one direction. Right-hand spiral
reamers have better chip-removal capabilities and for this reason are often used for
reaming deep blind holes. Straight-fluted chucking reamers are usually designed with
staggered flutes. The flutes are irregularly spaced around the circumference of the
reamer body to prevent setting up vibrations that will result in chattering. Chatter is
also reduced in some cases by using a reamer having a hand or spiral opposed to the
hand of the cut. Figure 4.140a shows a standard chucking reamer.

Shell reamers Shell reamers (basically machine chucking reamers without a shank)
are often referred to as hollow reamers because a tapered hole (1 mm per 100 mm)
through the centre allows the reamer to be held on a specially designed arbor.
They are available in sizes from 20 to 75 mm in diameter. The major advantage is that
several sizes of reamers can be used on one arbor. Only seven arbors are required to
fit all reamers from 20 to 75 mm. Arbors are made slightly larger at the large end of the
taper than the hole in the shell reamer to permit a driving fit with the reamer. The major
disadvantage is that the reamer assembly is substantially a two-piece assembly that
lacks the rigidity of a solid reamer. For this reason, a solid reamer is recommended for
reaming hard and tough materials. Figure 4.140b shows a shell reamer and mating
arbor.

Stub screw-machine reamers These reamers are designed primarily for use in
automatic screw machines, where short tools are required, but they can be used in
drill presses and hand-screw machines. They are particularly adapted to use in float-
ing holders, and a pinhole is provided in the shank for this applications. Figure 4.140c
shows a stub screw-machine reamer.

Expansion reamers Reamers of this type utilise a tapered pin at the end of the
reamer that causes expansion of the reamer body. Machine expansion reamers are
designed to maintain the initial size by compensating for wear on the cutting end.
When the reamer is worn, it can be expanded oversize and then reground to size.
The expansion pins of machine expansion reamers should never be loosened in an
attempt to use the reamers for a size smaller than that to which the tools were origi-
nally circle ground. Figure 4.140d shows an expansion machine reamer.
Expansion hand reamers are used to enlarge holes slightly in order to secure desired
fit. These reamers are made with an adjusting screw for expansion only and with an
undersized pilot to aid in alignment. They are ground with an entering taper to provide
easy starting. These reamers are designed to cut a small amount over nominal size
356 Tool Design

(a)

(b) DoALL

DoALL

(c)

(d)

(e)

(f)

(g) DoALL

DoALL
(h)

(i)

Fig. 4.140 Types of reamers: (a) Chucking reamer (b) Shell reamer and arbor (c) Stub
screw-machine reamer (d) Expansion machine reamer (e) Expansion hand reamer
(f) Adjustable hand reamer (g) Roughing hand taper reamer (h) Finishing hand
taper reamer (i) Carbide-tipped machine reamer [(a), (c) to (e), and (i): The
Standard Tool Company (b), (f) to (h): The DoAll Company]

for fitting purposes. The limit of expansion will vary from 0.15 to 0.3 mm, depending
upon the design of the tool. Figure 4.140e shows an expansion hand reamer.
It should be noted that neither machine nor hand expansion reamers are designed to
expand over a large range of sizes. Many expansion reamers are broken in attempting
Design of Cutting Tools 357

to expand them an amount beyond that for which they were designed. When a range
of sizes is necessary, adjustable reamers should be used.

Adjustable reamers This type of reamer is designed to permit setting any desired
diameter within the range of adjustment. This adjustment allows them to be set to a
special diameter, and when reamers wear, they can be set oversize, reground, and
resharpened. This reamer has removable, replaceable blades which fit into slots hav-
ing tapered bottoms. They are held in place by locking nuts or screws and advancing
an adjusting collar which advances each blade simultaneously. Figure 4.140f shows
an adjustable hand reamer.
The disadvantage of both expansion reamers and adjustable reamers is the lack of
necessary strength for heavy reaming operations. They also lack the accuracy for
extremely close tolerance work.

Taper reamers Reaming tapered holes is somewhat different from reaming straight
holes in that the taper reamer cuts on its periphery and more length of the cutting
edge is used as the reamer enters the hole. At the end of the operation, the cutting
edges are engaged in the cut through most of their length. A taper reamer therefore
serves as a roughing tool for heavy stock removal as well as a finishing tool. It is
constructed as heavy as possible for the reason. Taper reamers for machine reaming
utilise flutes with a left-hand spiral angle of 45°. The left-hand spiral (right-hand cut)
prevents the reamer from “sucking in” and still provides the efficiency of a shearing
cut. It does, however, increase the end pressure required to feed the reamer into the
work, necessitating the use of a machine tool.
In hand operations, two taper reamers are generally needed: one for roughing and
one for finishing. Roughing reamers are made undersize to allow stock for finish ream-
ing. The cutting edges of a roughing reamer are notched to form teeth that break
the chip and allow easier penetration of the workpiece material. Finishing reamers
are used for final finishing or for removing burrs or nicks in machine spindles. Figure
4.140g and h show both roughing and finishing hand taper reamers.

Carbide-tipped reamers Reamers of this type are well adapted to reaming highly
abrasive materials or for use on long production runs because of their wear-resisting
characteristics. They are well suited for reaming cast iron, where sand and scale are
encountered. Carbide reamers are available in the majority of standard machine-
reamer designs, including the expansion and adjustable types. Figure 4.140i shows a
typical carbide-tipped reamer.

Troubleshooting reaming problems Reaming is a finishing operation. Holes are


reamed to produce close tolerance dimensions and to improve the surface quality. A
reamer that produces a hole 25 mm oversize or a slight bell mouth at the beginning
of the hole will quickly attract the attention of the quality control department. The tool
designer may be called upon to correct the 25 mm oversize. In other words, what
may not be a problem in a drilling operation may be a large problem in a reaming
operation.
Reamers act differently in different materials. A given reamer may cut to size in one
material and cut oversize in another. The application or absence of a cutting fluid may
cause a reamer to act differently. Special reamer geometry may be required for certain
applications. Sometimes a reamer is ground undersize to produce desired results.
358 Tool Design

A great deal of experience is required on the part of the tool engineer to solve all
reaming problems efficiently. The following discussion is designed to give the begin-
ner an idea of the general problems that exist in a reaming operation. Years of practi-
cal experience will be required to make the beginner an authority.
One of the major problems in reaming operations is improper alignment with the
drilled hole in the workpiece. It is true that the reamer will generally follow the path of
the previously drilled hole, but this does not mean that it will cut equally on all sides
of the hole. When reaming on the drill press, the reamer is often forced into a partially
aligned hole in a manner to cause the workpiece to shift into alignment. The result
is a bell-mouthed hole and overloaded teeth on the reamer. The use of movable jigs
and bushings help to prevent bell-mouthed holes, but reamer damage may still occur
if the reamer is used as a locating device. Ideally, fixed jigs and bushings in proper
alignment should be used, but this is not always possible when multiples of holes
must be reamed.
The use of pilots may solve alignment problems, especially when the hole is relatively
long. The reamer should be guided from both sides of the work, as shown in Fig.
4.141. Piloted reamers of this type can be purchased through special order from most
of the major reamer manufacturers. The pilots should be grooved to allow cutting fluid
to enter and lubricate the bushing. The pilot acts as a chip cleaner.

Bushing

Workpiece

Piloted Bushing
reamer

Fig. 4.141 The use of piloted reamers and guide bushings

The above discussion stresses the importance of proper alignment for good reamer
performance. However, proper alignment is not always possible under actual plant
conditions because of worn spindles and worn reamer-holding devices. The result is
runout between the reamer and bushing (or drilled hole). When reaming with a rigid
jig, a condition of strain reversal exists within the reamer body that leads to premature
failure, along with alternate deflections of the reamer within the hole. The effect of
runout on sliding jigs is that the jig body shifts back and forth on the drll-press table
a distance equal to the amount of runout. A condition of this type cannot produce
satisfactory results. The most common solution is to use a floating reamer driver, as
Design of Cutting Tools 359

shown in Fig. 4.142. (An angular floating driver permits angular variation between
the axis of rotation and the axis of the reamer. A parallel floating driver permits side
movement of the reamer axis with that of the axis of rotation). A floating driver which
allows both parallel and angular float is best suited for general use, in that it adjusts
itself in all positions throughout the circle of rotation. The reamer is able to guide itself
into a bushing or previously drilled hole, where otherwise a bell-mouthed, tapered,
out-of-round, or oversize hole would be the result. Floating drivers can be obtained
commercially as shown in Fig. 4.143.

Spindle

Floating
driver

Parallel
float

Angular
float

Fig. 4.142 Movement of floating reamer drivers (The Metal Cutting Tool Institute)

Floating drivers can also be used to good advan-


tage when reaming with turret lathes or screw
machines. The axis of the reamer theoretically
should coincide with the axis of the machine
spindle, but this condition seldom exists in old
or worn machines. Angular or parallel misalign-
ments usually are present, and the result is bell-
mouthed and oversize holes. The use of floating Fig. 4.143 Commercial floating
drivers is a solution when simple corrections can- reamer driver (Erick-
not be made on the machine and toolholders. son Tool Company)
Misalignment errors do not always show on extremely loose machines because the
looseness itself may allow a certain amount of float. Another method of correcting
misalignment on machines where tools are not frequently changed is to insert a blank
bushing in the toolholder and rebore it from the spindle.
360 Tool Design

Chatter is another problem often encountered in reaming operations. Like any other
machining operations, the lack of rigidity promotes chatter when reaming. The use of
floating drivers certainly does not help to increase rigidity, and therefore their use may
be objectionable in some situations.
Chatter can ordinarily be eliminated by reducing the speed or by increasing the feed,
in that order. Reducing the lead angle clearance may also reduce chatter tendency
because less clearance prevents the cutting edges from digging into the workpiece.
Other remedies are to use unevenly spaced flutes to break up any tendency to sychro-
nise slippage and torsional deflection, to use spiral-fluted or piloted reamers, or to
chamfer the drilled hole before reaming.

4.13 TAPS
The standardisation carried out in thread forms and threading methods rarely makes it
necessary for the tool designer to design special taps. However, the design of special
tapping fixtures is quite common, and in order to accomplish this, the tool designer
must understand the construction and cutting action of taps. This knowledge is also
necessary when selecting taps. The same tap may act differently when used in differ-
ent materials. The size of a tapped hole depends upon the tapping equipment used,
alignment, speed of operation, material to be tapped. Size and condition of the hole,
and the type of cutting fluid used.

The tapping process The Metal Cutting Institute defines a tap as “a cylindrical or
conical thread-cutting tool with one or more cutting elements having threads of a
desired form on the periphery. By a combination of rotary and axial motion, the lead-
ing end cuts an internal thread while the tap derives its principal support from the
thread it produces.”
A tap is simply a hardened tool-steel screw with lengthwise grooves, called flutes,
milled or ground across the threads. The flutes form a series of teeth and provide chip
room along the entire length of the threaded portion of the tap. When the tap is turned
into a hole of proper diameter, the teeth cut into the wall of the hole and remove mate-
rial to form threads of the same pitch as the threads of the tap. The majority of the
work of a tap is done by the chamfer and the first full following thread.
Each tooth on the tap has rake and relief similar to that of other edged cutting tools.
Rake is determined by the position and shape of the cutting face of the tooth, as
shown in Fig. 4.144. The rake angle is determined by the angle at which the cutting-
face line intersects a line passing through the centre of the tap. Hook rake indicates
that the cutting face is concave in the area between the thread crests and the minor
diameter. The angle formed by a straight line passing from the thread crest through
the minor diameter in relation to the centreline of the tap determines the hook angle.
Relief is formed by using an eccentric grind behind the cutting edge. The machine-
screw sizes and small fractional sizes are usually ground radially or concentrically
with a radial area at the cutting edge followed by a relieved area at the heel of the
land. Still other taps, especially taper pipe taps, have a eccentric grind with the relief
extending to the cutting edge (see Fig. 4.145).
Design of Cutting Tools 361

Rake angle
Hook angle

Hook Rake

Fig. 4.144 Rake angles applied to taps

NO P. D. P. D. relief
relief to the Concentric
cutting edge margin
Cutting
Heel Eccentric
edge
relief

Concentric Eccentric Con-eccentric

Fig. 4.145 Methods of providing relief on taps

Nomenclature of tap elements Figure 4.146 shows a graphic illustration of tap


terms. They are further explained as follows:
Angle of thread The angle included between the sides of the thread, measured in
an axial plane.
Axis of tap The longitudinal central line through the tap.
Base of thread The bottom section of a thread; the greatest section between the two
adjacent roots.
Body Threaded and fluted part of tap.
Chamfer The tapered outside diameter at the front end of the threaded section.
Crest The top surface joining the two sides of a thread.
Cutting face The front part of the threaded section of the land.
Depth of thread The depth, in profile, is the distance between the top of crest and
the base or root of thread measured perpendicular to the axis of the tap.
External (male) Centre The cone-shaped end of the tap. It is incorporated for manu-
facturing purposes and usually at the threaded end of small taps only.
362 Tool Design

Fig. 4.146 Tap nomenclature

Flute The groove providing for the cutting faces of the threads or teeth, chip pas-
sage, and lubrication.
Heel The back part of the threaded section of the land.
Helix The curve formed on any cylinder, especially a right-circular cylinder, by a
right line in a plane that is wrapped round the cylinder, for example, an ordinary screw
thread.
Helix angle The angle made by the helix of the thread at the pitch diameter with a
plane perpendicular to the axis.
Hook The curved undercut of the cutting face of the land.
Internal (female) centre A small drilled and countersunk hole at the end of the tap,
necessary for manufacturing purposes.
Land The threaded web between flutes.
Point diameter The outside diameter at the front end of the chamfered portion.
Radial The straight cutting face of a land which, if continued, would pass through
the centre of the tap.
Rake The angle of the cutting face of the land in relation to a straight line from the
point of the cutting face to the axis.
Root The bottom surface joining the sides of two adjacent threads.
Shank The part behind the threaded and fluted section of tap.
Design of Cutting Tools 363

Side of thread The surface of the thread which connects the crest with the root.
Square The squared end of the tap shank.
Thread The cutting tooth of the tap which produces the thread in a tapped hole.
Thread relief (radial) A clearance providing a gradual decline in the major, pitch, and
minor diameters of the lands, back of the cutting face. This style of relief is applied
only to certain sizes and types of taps.

4.14 TAP CLASSIFICATION


Tap classification is generally based on construction. The solid tap, shell tap, expan-
sion tap, inserted chaser tap, and collapsible taps are a few examples. In this text,
only the solid hand tap will be discussed, as it is used in the great majority of tapping
operations. Hand tap is a general term denoting a specific style of tap of standardised
dimensions, having a cylindrical shank with a driving square. The hand tap was origi-
nally designed to be used by hand but now is almost exclusively in machines. The old
name is still used, however, and the beginner is often led to believe that hand taps
should be used for hand operations only.

Standard hand taps Standard hand taps are normally made with four straight flutes.
A set of standard hand taps refers to a set of three taps consisting of a taper, plug,
and bottoming tap. The amount of chamfer (number of tapered threads at the point of
the tap) determines whether the tap is classified as taper, plug, or bottoming. The plug
tap is most commonly used. Taper, plug, and bottoming taps are alike in every respect
except the number of chamfered threads at the point.
For the general run of tapping operations, standard four-flute hand taps have suf-
ficient flute space to permit chip passage. However, there are certain types of taping
jobs (such as tapping soft and stringy metals, deep holes tapped horizontally in screw
machines, or deep blind holes tapped vertically where the chips usually remain in the
flutes of the tap during the tapping operation) in which a tap with or three flutes may
be used, as fewer flutes provide more space for the passage of chips. Figure 4.147
shows a set of standard hand taps.

Bath Bath

(a) (b)

Bath

(c)

Fig. 4.147 A set of standard taps: (a) Taper tap (b) Plug tap (c) Bottoming tap (John Bath
& Company)

Spiral-point hand taps This tap, often called a gun tap was developed to overcome
problems caused by chips accumulating in a hole while threading and also causing
damage to the tap on reversal. The spiral point produces a shearing action in use that
364 Tool Design

projects the chips ahead of the tap. In open or through holes, where chips can exist
ahead of the tap, this tool is ideal. It can also be used in blind holes providing there
is sufficient depth beyond the threads to accommodate chips. Because the flutes are
not needed for chip passage, they can be made shallower than in standard hand taps,
resulting in a thicker recommended for through-hole machine tapping in most materi-
als (see Fig. 4.148).

Sossner

Fig. 4.148 Spiral-point hand tap (Helicoil Cutting Tool Division)

Spiral-fluted hand taps The spiral-fluted taps (sometimes called helical-fluted taps)
have helical flutes at the same hand of flute lead as the hand of cut and cause the
chips to flow toward the shank of the tap. They are especially useful when tapping
blind holes in ductile materials which produce long curling chips or where the cutting
action is interrupted by slots, keyways, or other gaps. Chips are guided out by the lift-
ing action of the spiral flutes and a relatively chip-free tapped hole is produced. The
fine chips remaining in the tapped hole are easily removed. The shearing cut also
reduces strain on the tap and power requirements for tapping the hole. Figure 4.149
shows the spiral-fluted tap.

Besly

Fig. 4.149 Spiral-fluted tap (Besley-Welles Corporation)

Troubleshooting tapping problems Tapping problems are usually confined to tap


breakage, oversize and undersize threads, rough or torn threads, wavy threads, and
material buildup on teeth. Table 4.24 gives a summary of causes and prevention of
tapping problems.
Tap breakage is probably the most common difficulty encountered in tapping. Hitting
the bottom of the hole, incorrect tap style, incorrect hole size, chips packing in
fluted, misalignment, and insufficient lubrication are major causes of tap breakage.
Sometimes tap breakage will occur when the tap is reversed. Some materials tend
to flow back and close in around the tap after the thread is cut. The result is seizure
and tap breakage when the direction of operation is reversed. Materials that form a
continuous chip may also cause problems when the tap is reversed because the heel
of each tooth must cut through the chip that remains attached when the following
tooth stop cutting.
Tap breakage is greater when tapping the newer alloys used in the aerospace indus-
tries, especially when a tap drill is used to produce the conventional 75 per cent
thread height. (The thread height is determined by the size of the tap drill. If the tap
drill is the same as the minor diameter, the height of the thread is less than 100 per
cent height; if the tap drill s larger than the minor diameter, the height of the thread is
Table 4.24 Sumarry of causes and prevention of tapping problems
Problem
Blind Brea- Wear Load Torn Wavy Under Over Bell Cause Prevention
ing kage ing threads threads size size mouth

Chips: Heavy-duty tapping lubricant, use larger flutes,


× × × × × × Packed in flutes spiral point, or spiral flute
× Packed bottom Use spiral flute, M-boss or less % thread
Tap:
× × × × × × Dull Sharpen tap
× × Chamfer length Lengthen chamfer evenly
× Chamfer relief Reduce
× × Flute design Use fewer flutes, internal thread, spiral point or spiral flute
× × × Thread relief Reduce relief; increase for binding
× Pitch diam small Higher H limit
× × × Pitch diam large Lower H limit
× × Hook angle Increase for torn threads; reduce for high hardness
× Land width Reduce
Drilled hole:
× × × Small Check drill size
× × × Work-hardened Sharpen drill
× × Out of round Check drilling setup
× Torn surface Sharpen drill

Contd...
Design of Cutting Tools 365
Problem
Blind Brea- Wear Load Torn Wavy Under Over Bell Cause Prevention
ing kage ing threads threads size size mouth
366 Tool Design

Tapping machine:
× Wrong size Use proper machine for tap size
× Slippage Check power transmission
× × × Excessive feed pressure Adjust pressure, consider lead screw
× × × × Excessive speed Adjust speed
× × × × × Slow speed Adjust speed
× × Excessive spindle float Reduce float
Worn holder or loose
× × × spindle Replace or repair
× × × Misalignment Align spindle, fixture, and work
× × × Hard material Heavy-duty tapping lubricant, nitride or both.
Reduce hook and speed, increase relief
× × × × Welding Heavy-duty tapping lubricant
× × × × × × Lubrication Increase flow; check for proper type

SOURCE: Helicoil Cutting Tool Division.


Design of Cutting Tools 367

less than 100 per cent.) The pocket-size tap-drill size charts distributed by drill com-
panies are all 75 per cent thread height. They have been so widely distributed that
it has become common belief that all threads should have 75 per cent thread height
for adequate strength. However, a nut with only 50 per cent thread height will break
a bolt before it will strip the threads. Furthermore, thread-strength tests clearly show
that any increase in thread height over 60 per cent for the tapped member does not
increase the static strength of a threaded fastening. On the other hand, the torque
required to drive the tap is double when the percentage of thread height is increased
to 80 per cent. Thus, the chances of tap breakage are increased without securing
any added strength when the thread height is increased. Reducing the thread height
where tap breakage is encountered may eliminate the problem and yet still provide
sufficient strength.

Tapping speeds The feed and speed of taps are not independent of each other
because the feed per revolution is governed by the lead of the thread.
The selection of tapping speeds depends upon many variables such as the material
being tapped, the cutting fluid, tap style, type of hole (through or blind), depth of hole,
percentage of thread, thread pitch, condition of equipment, method of tapping, and
tap chamfer. Taps with long chamfers may be run at higher speeds because the long
chamfer reduces the chip load per tooth. Coarse thread taps and taps with fewer
flutes must be run slower for the same reason. Table 4.25 lists speeds proved satis-
factory under average conditions to serve as a starting point. The best speed is then
determined by experiment on the job.

Table 4.25 Speeds and lubricant for machine tapping


Material Lubricant Tapping speed,
m/min

Aluminium Kerosine and lard oil 30


Bakelite Air blast 25
Brass Soluble or light = 1 base oil 40
Cast iron Dry or soluble oil 25
Steel, mild Soluble or sulphur-base oil 18
Medium alloy Sulphur-base oil 12
Stainless (300 series) Sulphur-base oil 06
Zinc diecast Soluble oil 25

4.15 THE SELECTION OF CARBIDE CUTTING


TOOLS
4.15.1 Carbide Tools
Cemented carbides, products of powder metallurgy, are composed of various combi-
nations of tungsten carbide, titanium carbide, tantalum carbide, and cobalt. Powered
carbides are mixed in various proportions and are ball-milled with cobalt powder. After
ball milling, the mixture is pressed into the desired shape and sintered at high temper-
atures in a controlled-atmosphere furnace. During the sintering operation, the cobalt
368 Tool Design

melts and essentially bonds the carbide grains together to form the final cemented-
carbide material. The various compositions determine the properties and grade of the
finished cemented-carbide cutting tool.
Manufacturers of cemented carbide cutting tools offer several hundred different
grades in a wide variety of compositions. To simplify the selection of the many avail-
able grades for a particular machining operation, an attempt has been made to group
the grades by methods of application, as shown in Table 4.26. The grouping system
at best is a compromise because of the vast number of variables in each manufac-
turer’s product. The composition of each manufacturer’s grade will differ although
they may all fall under the same application grade shown in the comparison chart.
Also, the applications shown in the chart may not have the same meaning to different
people. For example, a roughing cut in a plant producing automotive parts would not
be the same as a roughing cut in a plant producing products for the shipyards. The
grade-comparison charts give only a general idea of a given manufacturer’s grades
and should be used only as a starting point.
No great problem is encountered when selecting grades for machining soft steels
or cast irons under satisfactory conditions. In this case, a grade could probably be
selected according to comparison-chart recommendations. However, when machin-
ing alloys under difficult machining conditions, the selection of grade by comparison
chart may not work. The recommended grade may fail prematurely. It becomes neces-
sary to select another grade based upon how the tool failed. For examples, tool failure
by edge wear would indicate the need for a grade more resistant to abrasion, while
failure by chipping the cutting edge could indicate the need for a stronger grade. In
order to make an intelligent selection, the tool engineer should understand the effect
that changing the variables (carbides and cobalt) has upon the physical properties of
the cutting tool.
Tungsten carbide cutting tools containing only tungsten carbide and cobalt with no
other elements added are referred to as the straight grades. The only variables in this
case are the grain size of the tungsten carbide particles and the percentage of cobalt.
Any change of these variables affects the hardness and transverse rupture strength
(toughness) of the carbide insert. In turn, any reduction of hardness would reduce the
abrasion resistance of the tool, and a reduction in toughness would result in a more
brittle tool. In other words, the tougher the grade, the lower its wear resistance and
vice versa.
The effect of the variables in the straight carbide grades is as follows: (i) an increase
in grain size lowers the hardness, (ii) an increase in grain size increases the trans-
verse rupture strength (toughness), (iii) an increase in the cobalt content lowers the
hardness, (iv) an increase in the cobalt content lowers the hardness, (v) an increase
in the cobalt content increases the transverse rupture strength.
This shows that maximum wear resistance is obtained from using a small grain size
and cobalt content. These grades are listed in general grade classification charts as
C-1 to C-4; however, this does not mean the C-2 grade at one company contains the
same percentages as the C-2 grade at another.
The straight carbide grades are used primarily for machining materials that produce
tool wear by abrasion. The chip produced when using these materials is generally
of low strength, brittle, and segmented. Cast irons, nonmetallics, and nonferrous
alloys are typical examples. The straight carbide grades are often referred to as
Table 4.26 Carbide grade chart
Listings do not necessarily imply equivalency of various manufacturer’s grades.

Machining applications
Cast-iron, nonferrous, and nonmetallic materials Steel and steel alloys

Carbide C-1 C-2 C-3 C-4 C-5 C-6 C-7 C-8


manu- Roughing General- Finishing Precision Roughing General Finishing Precision
facturers purpose finishing purpose finishing

Adamas B A, AM, PWX, AA AAA DD, 5X, 6X, D 7X, C, CC,


PWX 434 548, Titan 80*
Titan 80*
Amcarb ... D15, D13 ... ... ... ... ... ...
Besly-Welles B101 B106, B108 B211 B109, B102 B103, B207
B168 B221, B104, B365*
B205,
B245
Carboloy 44A 883, 860 883, 905, 999, 895, 370, 78B 370, 78B, 350, 78, 320
895 320 78, 350 320
Carmet CA-3 CA-4, CA-7 CA-8 CA-610, CA-606, CA-711 CA-704
CA-443 CA-740 CA-720
Coromant H20 H13, H20, H1P H05 S6, S4 S2 S1P F02*
H1P
Firth-Loach FA-5 FA-6 FA-7 FA-8 FT-3, FT-5, FT-6 FT-7,
FT-4, FT-62 FT-62 FT-72*
FT-5
Firth Sterling H HA, H-23 HE HF T04, TXH, T22, T31,
NTA T22 TXL WF*
Futurmill ... DMC21 ... ... DMC30, DMC32 DMC35 ...
DMC32
Kennemetal K1 K6, K68, K8 K11 KM, K21, K2S, K45, K7H,
Design of Cutting Tools 369
Machining applications
Cast-iron, nonferrous, and nonmetallic materials Steel and steel alloys

Carbide C-1 C-2 C-3 C-4 C-5 C-6 C-7 C-8


manu- Roughing General- Finishing Precision Roughing General Finishing Precision
370 Tool Design

facturers purpose finishing purpose finishing

C8735, K2S K3H, K5H, K165*


K68 K4H, K7H,
Multimetals 0M1 0M2 0M3 0M4 4M5 ... ... ...
New comer N10 N20 N30 N40 N50 N60 N70, N80,
NM-93* NM-93*
NM-95
Sintercast Ferro- Ferro- ... ... Ferro- Ferro- ... ...
Tic J Tic J Tic J Tic J
Speedicut mitia A B C C TA10, TTA TE TE
TA5
Talide C-89 C-91 C-93 C-95 S-880 S-901 S-92 S-94
S-900
Tungsten alloy 9 9H 9C 9B 11T, 9S, 9S, 10T, 8T, 5S 5S
10T 5S
Unimet U10 U20 U30 U40 U53 U53, U70, U73,
U60 U73 U80,
U88*
Valenite VC-1 VC-2, VC-3 VC-4 VC-125, VC-125, VC-7 VC-8
VC-22, VC-55 VC-6 VC-83*
VC-28 VC-85*
VR/Wesson 2A-68, 2A-5 2A-7 VR-52, WS, VR-75 VR-73, HV,
VR-54 VR-54 2A-7, VR-77, WM WH, HV, VR-733
VR-65* VR-89, VR-65* VR-65*
VR-75
Walmet WA-141, WA-2, WA-35, WA-4 WA-66, WA-5, WA-147, WA-8
Machining applications
Cast-iron, nonferrous, and nonmetallic materials Steel and steel alloys

Carbide C-1 C-2 C-3 C-4 C-5 C-6 C-7 C-8


manu- Roughing General- Finishing Precision Roughing General Finishing Precision
facturers purpose finishing purpose finishing

WA-1, WA-63, WA-3 WA-5 WA-6 WA-7


WA-159 WA-149
Wendt-Sonis CQ12 CQ2 CQ3 CQ4 CY12, CY16, CY14, CY31,
CY16 CY5 CY2, T18*
T18*
Wickaloy N H HH HHH X7A, X7 G8 GX FX
Willey’s E8 E6 E5 E3 945, 8A, 8A 606, 6A 6AX,
10A 509
*Grade containing more than 50% titanium carbide.
SOURCE: “Machining Data Handbook,” Metcut Research Associates, Inc.
Wear and impact implications
Wear surface Impact Miscellaneous
Carbide C-9 C-10 C-11 C-12 C-13 C-14 C-15 C-15A C-16 C-17 C-18 C-19
manufacturers No shock Light Heavy Light Medium Heavy Light- Heavy- Rock- Cold Wear at Radioactive
shock shock cut hot cut hot bits header elevated shielding,
flash- flash- dies temp and/ counter
weld weld or resis- balances,
removal removal tance to and kinetic
chemical applications
reaction
Adamas A B BB BB HD15 HD25 434, 474 474, GG 569, 783 HD25
502 HD 257
Amcarb D13, D30, D50, D50, D65, D80, D43 D55 DA120 D60,
D10, D5 D20, D40 D41A D50, D75, D70,
D13 D40 D41 D41A D80
Design of Cutting Tools 371

Besly-Welles
Wear and impact implications
Wear surface Impact Miscellaneous
Carbide C-9 C-10 C-11 C-12 C-13 C-14 C-15 C-15A C-16 C-17 C-18 C-19
manufacturers No shock Light Heavy Light Medium Heavy Light- Heavy- Rock- Cold Wear at Radioactive
shock shock cut hot cut hot bits header elevated shielding,
372 Tool Design

flash- flash- dies temp and/ counter


weld weld or resis- balances,
removal removal tance to and kinetic
chemical applications
reaction

Carboloy 44A, 883 44A, 779 55A, 55A, 779, 55B, 190 77B 77B 90, 241, 190 608, HM1,
895 55B 115 55A 115 4237 HM3
Carmet CA-4 CA-12 CA-10 CA-214 CA-11 CA-225 ... ... CA-210 CA-225, CA-815
CA-220
Coromant H05, H20, H35, H35, G4 G5 S2 S6 ... G5
H20 H25 H45 H45
Firth-Loach FA-6 FA-5 FA-3 FB-5 FB-3 FT-8 FT-9 FM-3, FH-2,
FM-5 FH-3,
FH-4
Firth Sterling HA H HC DC2, DCX DC3, T89 T75 MPD2 ND25 ... Firth,
DC1 DC4 heavy
metal
Futurmill
Kennametal K6, K95, 3109, 3109, 3109, K90, KM, 3365 3411, K90A, K151A, W10,
K96, K6, K92, K94, K92, K91, K162B 3047 K90, K162B, W2
K68 K96 K94 K96 K94 3109 K91 K601,
K701,
K801
Multimetal OM3 OM2 IM2 OM1 1M12 1M13 ... ... 1M13
Newcorner N20 N10
Sintercast Ferro- Ferro Ferro- Ferro- Ferro- Ferro- ... ... ... Ferro- Ferro-
Tic C Tic C Tic C Tic C Tic C Tic C Tic C Tic C
Wear and impact implications
Wear surface Impact Miscellaneous
Carbide C-9 C-10 C-11 C-12 C-13 C-14 C-15 C-15A C-16 C-17 C-18 C-19
manufacturers No shock Light Heavy Light Medium Heavy Light- Heavy- Rock- Cold Wear at Radioactive
shock shock cut hot cut hot bits header elevated shielding,
flash- flash- dies temp and/ counter
weld weld or resis- balances,
removal removal tance to and kinetic
chemical applications
reaction
S-45 &
S-55
Speedicut mitia B A RD-7 H M XT TA5 TA10 PSM RD7
Talide C-99 C-88 C-80 C-85 C-80 C-75 CT-28 S-880 C-88-X C75 CR83 W
C-90-X CT50 Alloy
Tungsten alloy 9 9M 9A 15C 9A20 9A30 16G 11C 25H,
9H 9 9A15 9A15 15C 9A25 30H,
20H
Unimet U20 U10 U110 U130 U140
Valenite VC-9 VC-10 VC-11 VC-12 VC-13 VC-14 VC-125 VC-55 VC-11 VC-14
VR/Wesson 2A-68, 2A-68, 2A-1, VR-13 VR-14 VR-15 VR-89 VR-87 VR-13,
2A-5 2A-6 2A-3 VR-14
Walmet WA-2 WA-10 WA-11, WA-12 WA-13 WA-140, WA-5 WA-64 WA-11, WA-140, WA-200 Waltung
WA-138, WA-14 WA-12 WA-14
WA-134
Wendt-Soins CQ2 CQ12 CQ14 CQ15 CQ13 CQ16 CX3 CQ13
CY4
Wickaloy N, HW D8, D15, D13 D20 D25 X7 X7 FBS, D25
D13 D20 FBT
FBY
Willey’s E6, E8 E9 E16 W12C, E16 E25 ... ... E9 E25
Design of Cutting Tools 373

E8
374 Tool Design

cast-iron grades, but this may be misleading in the sense that some types of mal-
leable cast irons that exhibit the properties of steel are better machined with steel-
cutting grades.
The straight carbide grade selected for a particular machining job should be just strong
enough to handle the cut without chipping or breaking. Usually, the first selection is
an estimate, with a trail cut being made to determine the effect of the work material
on the carbide grade. If the cutting edge shows signs of chipping, a tougher grade
(more cobalt) is required. On the other hand, if the flank of the cutting edge shows
excessive wear, a more wear-resistant grade is indicated (more tungsten carbide or
smaller grain size). It is often hard to determine which of the straight grades is harder
or tougher, but speaking in broad terms the C-1 grade is softer and tougher while the
C-4 grade is harder and more brittle. Some companies give the percentages of cobalt
and tungsten carbide in their catalogues, which makes it easier for the person who
understands the effect of varying the percentages to choose the type he wants.
If a crater develops in the top of the insert where the chip flows over it, the selection
of a grade other than the straight carbide grades in indicated. Most steels and highly
alloyed cast irons produce wear by cratering in straight carbide grades. High tempera-
tures and heavy pressures produced by a strong continuous chip create a tendency
for the chip to weld to microscopic particles of the carbide and “wash away” with the
chip. To combat this condition, titanium carbide is added to the basic tungsten carbide-
cobalt compositions. The titanium carbide lowers the tendency of the chip to weld
to the tool because the adhesion temperature between steel and titanium carbide
is higher than that between steel and tungsten carbide. An increase in the titanium
carbide content increases the resistance to cratering; however, it also decreases the
toughness and abrasive wear resistance. For this reason straight tungsten carbide
grades should always be used unless the crater-resistant property is needed. In any
case, the titanium carbide content should be kept as low as possible to still do the
job.
Tantalum carbide is sometimes added to tungsten carbide-cobalt materials to obtain
certain desired properties. The addition of tantalum carbide will reduce cratering, but
no more so than titanium carbide. It also lowers the strength about the same degree
as titanium carbide. Since tantalum carbide is much more expensive than titanium
carbide, it is not added to the prime purpose of reducing cratering. Additions of tan-
talum carbide do not effect the abrasive wear resistance as titanium carbide does;
therefore, it provides a means of improving cratering resistance without sacrificing
abrasive wear resistance. Tantalum carbide also combats softening and deformations
in operations where extreme temperatures are produced.
The titanium carbide and tantalum carbide grades are often referred to as steel-cutting
grades or crater-resistant grades. They are listed in general classification charts as
C-5 to C-8. Here again, general classification charts may be deceiving because many
of the stainless steels are machined with straight carbide grades with better results.

Summary

Metal cutting tools form an important aspect of tooling applications. This chapter sys-
tematically addressed the geometry, theory, design and related aspects of cutting tool
used in metal cutting.
Design of Cutting Tools 375

• Even with all of the sophisticated equipment and techniques used in today’s
modern industry, the basic mechanics of forming a chip remain the same. As
the cutting tool engages the workpiece, the material directly ahead of the tool
is sheared and deformed under tremendous pressure. The deformed material
then seeks to relieve its stressed condition by fracturing and flowing into the
space above the tool in the form of a chip.
• Different types of chips are formed during metal cutting depending upon the
work material, cutting material and the operating parameters selected. There
are three are three basic types of chips that are formed during metal cutting;
they are, discontinuous chips, continuous chips, and BUE chips.
• The two basic methods of metal cutting are the orthogonal and oblique cutting.
Orthogonal cutting takes place when the cutting face of the tool is perpen-
dicular to the line of action of the tool. If the cutting face is inclined at an angle
less then 90° to the line of action of the tool, the cutting action is known as
oblique.
• Analysis of metal cutting based on oblique cutting is complex. Therefore, the
preferred approach is orthogonal cutting for metal cutting analysis. Orthogonal
cutting is only an approximation of the actual scenario.
• Cutting is a process of extensive stresses and plastic deformations. The high
compressive and frictional contact stresses on the tool face result in a sub-
stantial cutting force. Knowledge of the cutting forces is essential for the fol-
lowing reasons: proper design of the cutting tools, proper design of the fixtures
used to hold the workpiece and cutting tool, calculation of the machine tool
power, selection of the cutting conditions to avoid an excessive distortion of
workpiece.
• In orthogonal cutting there are two component of forces that are acting on the
tool, cutting force and thrust or feed force. These two forces can be measured
using tool force dynamometer.
• During metal cutting, the chip is considered to held in equilibrium under the
action of frictional and shear force system.
• Merchant’s circle diagram can be conveniently used to determine the relation
between the various forces and angles.
• A practical approach for finding the power requirements in metal cutting is by
finding the specific energy of the material. Specific energy is defined as the
total energy per unit volume of material removed.
• The ease of machining a metal can be determined form machinability ratings.
Machinability rating is the ration of the specific cutting speed for certain tool life
of the test material to the specific cutting speed for same tool life of a standard
material.
• Cutting fluids, sometimes referred to as lubricants or coolants are liquids and
gases applied to the tool and work piece to assist the cutting operation. Their
main functions are, to cool the tool, to cool the workpiece, lubricate and reduce
friction, improve surface finish, protect the finished surface from corrosion and
wash the chips away from the tool.

Questions
1. Why has the tool designer’s job been simplified in recent years insofar as cut-
ting tools are connected?
376 Tool Design

2. What are the basic requirements of a cutting tool?


3. Explain the progressive formation of a metal chip during a machining
operation.
4. What terms are commonly used to define the shape of a metal cutting tool?
5. Why do positive rake angles have greater cutting efficiency from the standpoint
of power requirements and cutting forces?
6. What are the limitations of high positive rake angles?
7. What are the advantages of negative relief angles?
8. What is the purpose of relief angles?
9. When is it permissible to have negative relief angles?
10. Know and understand the approximate equivalents of relief and rake angles for
common edged cutting tools.
11. In chip formation, what is meant by the shear plane?
12. What is the difference between shear plane and shear angle?
13. How does the process of chip formation contribute to vibration, or chatter of the
cutting tool?
14. Name and describe the three basic types of chips.
15. What determines the type of chip?
16. What causes the BUE chip?
17. Why is the BUE chip associated with ductile materials?
18. Why does a BUE chip produce a poor surface finish?
19. Generally speaking, what relative cutting speed produces a BUE chip?
20. Name the design and machining practises that help to reduce the built-up edge
on the BUE chip.
21. Why is a large shear angle desirable?
22. What is meant by the cutting ratio of a chip?
23. Name and describe the three basic force components in oblique cutting.
24. Why does the tangential force constitute approximately 99 per cent of the total
power required by the cutting tool?
25. When is it necessary to keep the radial force to a minimum?
26. Why is the cutting-tool dynamometer one of the most reliable methods of mea-
suring cutting forces?
27. Why is it important to know the horsepower required at the motor of the machine
tool for a specific set of cutting conditions?
28. What is meant by power?
29. What effect does cutting speed have on cutting forces within the normal cutting
range?
30. Why is it possible to remove metal more efficiently at heavy feeds than heavy
depth of cut?
31. Why should tools with negative rake angles be operated at higher cutting
speeds?
32. What are the factors in the work material that help to determine cutting forces
and power required?
33. What is meant by machinability?
34. What material is used as the machinability index for steels?
Design of Cutting Tools 377

35. Why do low-carbon steels with less than about 0.15 per cent carbon machine
poorly?
36. Why does the increase of carbon over 0.20 per cent decrease the machinabil-
ity of steel?
37. What causes the cratering type of tool wear?
38. How is flank wear designated?
39. What determines how wide a wear land can be allowed before the tool should
be resharpened?
40. Why is flank wear used as the criterion of tool-life studies?
41. Why is it possible to predict flank wear after a wear rate has been established
through tool-life studies?
42. What factor is recognised as contributing to tool wear more than any other?
43. What are the two main areas of heat generation during chip formation?
Explain.
44. When a large amount of metal is to be removed and temperature is a prime
consideration, why is it better to increase the feed rather than speed?
45. How is the cutting speed of a given material designated?
46. Why is the feed of a reamer higher than that for drilling?
47. Why is milling feed determined by the amount of allowable chip load per cuter
tooth?
48. How do slower speeds better utilise available power?
49. What are three main factors that determine surface roughness during a machin-
ing operation?
50. Why does too large a nose radius tend to induce chatter?
51. Why will higher cutting speeds sometimes improve surface finish?
52. What is the difference between forced and self-excited vibrations?
53. Why is a 30 mm milling-machine arbor, twice as stiff (rigid) as a 25 mm arbor
of the same length?
54. If the overhang of a 16 mm diameter boring bar is increased from 50 to 25 mm,
how much more will it deflect under a given load?
55. Will changing the tool material of a boring bar to a better grade of steel or hard-
ening to a higher hardness improve the rigidity of the bar? Why?
56. In view of the above questions, why are boring bars heat-treated?
57. What is the main purpose of cutting fluids?
58. What are the broad general classifications of cutting fluids?
59. What is hydrodynamic lubrication?
60. What is boundary lubrication?
61. With regard to cutting fluids, why is boundary lubrication more important than
hydrodynamic lubrication?
62. Why is pure petroleum oil not suitable as a cutting fluid?
63. What are the advantages of fatty mineral oils?
64. What is meant by an active or inactive cutting fluid?
65. How does the addition of sulfur to a mineral oil provide boundary lubrication?
66. Why is a chlorine additive sometimes introduced into cutting oils in combina-
tion with sulfur?
378 Tool Design

67. What is an emulsifiable or soluble mineral oil?


68. What is the main advantage of emulsifiable cutting oils?
69. Specify a suitable cutting fluid for each of the following machining operations
and materials: (a) Broaching stainless steel, (b) Milling aluminium, (c) Gear-
cutting high-alloy steels, (d) Threading low-carbon steel, (e) Grinding alloy
steel, (f) Threading brass, (g) Automatic screw-machining of magnesium.
70. Why are emulsions not recommended for machining magnesium?
71. When machining with carbide or oxide cutting tools, why may an insufficient or
intermittent flow of coolant be detrimental to tool life?
72. What would be the best way to apply cutting fluids to tool and cutter grinding
operations?
73. What is the advantage of the pressure-fed type of mist generator over the aspi-
rator type?
74. How is the size of a single-point tool designated?
75. Identify the following elements of a single-point tool: (a) Shank, (b) Flank,
(c) Heel, (d) Face, (e) Cutting, (f) Nose, (g) Side cutting-edge angle, (h) End
cutting-edge angle, (i) Side relief angle, (j) End relief angle, (k) Side clearance
angle, (l) End clearance angle, (m) Back-rake angle, and (n) Side-rake angle.
76. What is a tool signature?
77. Why is a SCEA of 15 to 30° an excellent choice for general machining (unless
determined by the workpiece shape)?
78. Why should a ECEA of 4° or less be avoided?
79. What is the result when too little relief is given to the cutting tool?
80. What is the difference between clearance angles and relief angles?
81. What angles determine the flow of the chip across the face of the tool?
82. Why are negative rake angles necessary when taking interrupted cuts with
carbide tools?
83. What tool materials are usually supplied in solid-type single-point tools?
84. Why do solid-type tools held in the toolholders shown in Fig. 4.67 require addi-
tional relief of from 15 to 20°?
85. What are the three basic methods generally used for brazing carbide tips to
tool shanks?
86. What is the advantage of induction heating when brazing carbide tips?
87. What are the advantages of brazed carbide tools?
88. What problems are encountered when carbide is brazed to steel?
89. How are on-end carbide inserts resharpened?
90. What is the major advantage of negative rake inserts used in throwaway insert-
type tools?
91. What is the purpose of the carbide seat provided to support the throwaway
insert?
92. Why is there a tendency for the insert to shift in the pocket of a throwaway
toolholder during tracing operations?
93. What are the advantages and disadvantages of the screw-, bridge-, and pin-
type clamping mechanisms of toolholders?
94. What is the advantage and the disadvantage of the angular ground-in chip
breaker?
Design of Cutting Tools 379

95. What are the dimensions for a step-type chip breaker for a 5 mm depth of cut
and 0.4 mm feed?
96. What provisions are made for adjustment of mechanical chip breakers to meet
the demands of different cutting conditions?
97. Why does the pressed-in chip breaker produce acceptable chips over a wide
range of materials, feeds, speeds and depths of cut?
98. Why do problems occur in boring that are generally not associated with exter-
nal turning?
99. When is it advantageous to set the boring bar above the centreline of the work-
piece? Why?
100. Referring to Question 99, what must be done to the boring-tool geometry when
the boring bar is raised above centre?
101. A boring tool is set 1.6 mm above the centreline of a 45 mm bore in mal-
leable iron. What should the rake and relief angles be for an on-centre cutting
action?
102. Why is chip formation in milling more complicated than in single-point
turning?
103. Generally speaking, why is climb milling more desirable than conventional
milling?
104. Why is it generally impossible to do climb milling on older machines?
105. How is backlash eliminated on modern milling machines?
106. Why is it important that the diameter of a face-milling cutter be wider than the
workpiece?
107. What is a profile-relieved milling cutter?
108. What is the major advantage of a form-relief cutter?
109. Be able to identify standard milling cutters.
110. Be familiar with the nomenclature of milling-cutter elements.
111. When selecting a milling cutter, why is it important to keep the cutter diameter
as small as possible?
112. What is the general rule for the number of teeth when selecting a milling cutter?
Why?
113. Generally speaking, why should milling cutters with straight flutes be
avoided?
114. When is it necessary to grind a secondary relief on a milling cutter?
115. What is the advantage of eccentric relief on small milling cutters?
116. Why is it preferable to use face-milling cutters with a chamfer sufficiently wide
for cutting to be confined to the cutting edge along the chamfer when face-
milling a plane surface?
117. What is the effect on feed when a wide chamfer is used on a face-milling
cutter?
118. What is the advantage of a wedge-behind-the-insert design of indexable-insert
face-milling cutters?
119. What is the purpose of a wiper insert used in indexable insert face-milling
cutters?
120. Why is chip formation produced in a drilling operation extremely complex?
121. Why is the chisel edge a problem area during a drilling operation?
380 Tool Design

122. What is the major purpose of the helical flute of a twist drill?
123. How is it possible for a drill to unwind during a drilling operation? How is this
tendency reduced?
124. What is the amount that a properly sharpened 13 mm drill will cut oversize?
A 20 mm drill? A 25 mm drill?
125. What is the major advantage of screw-machine-length drills?
126. What is the advantage of high-helix drills? Low-helix drills?
127. How is the effective rake angle changed on a standard drill when the drill geom-
etry is fixed by the manufacturer?
128. Why will increasing the feed rate of a drilling operation sometimes cause the
chip to break?
129. What should be done to correct for drill failure when the outer corners of the
drill have been wiped off during the drilling operation?
130. What are the possible causes of oversize holes when drilling?
131. What causes chipped cutting lips (along the entire length) during drilling?
132. What causes a drill to split up the web?
133. How much material should be removed when reaming a 6 mm hole?
134. Why is reaming speed slower than drilling speed?
135. What is the major difference between a hand and machine reamer?
136. What is the difference between a rose chucking reamer and a fluted chucking
reamer?
137. What is the advantage of a rose reamer? A fluted reamer?
138. What is the standard angle of chucking reamers?
139. When should reamers with a left-hand spiral be used?
140. Why are straight-fluted chucking reamers designed with staggered flutes?
141. What is the major difference between an expansion reamer and an adjustable
reamer?
142. What is probably the major problem encountered in reaming operations?
143. How can improper alignment be corrected when reaming on machines with
worn spindles and worn reamer-holding devices?
144. What are the major advantages of a spiral-point hand tap?
145. What are the major advantages of a spiral-fluted hand tap?
146. Why does tap breakage sometimes occur when tap is reversed?
147. What determines the thread height of the thread in a tapped hole?
148. How does reducing the thread height help to solve tap-breakage problems?
149. What is meant by the straight grades of tungsten carbide cutting tools?
150. What are the variables and their effect in the straight carbide grades?
151. What materials are machined with the straight carbide grades?
152. What is the effect of adding titanium carbide to the basic carbide-cobalt com-
position of carbide cutting tools?
153. What is the effect of adding tantalum carbide to the basic carbide-cobalt com-
position of carbide cutting tools?
154. When are tantalum carbide grades used in place of titanium carbide grades?
Design of Cutting Tools 381

Problems
1. An orthogonal cut of 3 mm depth produces a chip thickness of 3.8 mm.
(a) What is the cutting ratio? (b) What cutting action and quality of surface finish
can be expected?
2. Referring to Prob. 1, what would be the shear angle if the back-rake angle were
10°?
3. An experimental orthogonal cut 0.25 mm. deep made with a tool containing
a back-rake angle of 10° produces a chip 0.6 mm thick. The shear angle was
measured by use of a toolmaker’s microscope and was determined to be
24° (tool was stopped in cut and slip plane was measured). Is this shear angle
theoretically correct?
4. Determine the power required at the motor for a turning operation with the
following conditions:
Machine: geared-head lathe
Workpiece material: AISI 1055 steel, BHN 265
Cutting speed: 92 m/min
Feed: 0.65 mm/rev
Depth cut: 6.4 mm
Cutting tool: carbide, steel-cutting grade with 15° SCEA
5. It is desired to machine rolled aluminium on a 2.24 kW engine lathe under
the following conditions. Does the lathe have the necessary power for the
proposed machine operation?
Machine: V-belt driven lathe (2-belt)
Workpiece material: rolled aluminium
Cutting speed: 120 m/min
Feed: 0.25 mm/rev
Depth cut: 3 mm
Cutting tool: high-speed steel with 30° SCEA
6. Find the power requirements for a face-milling operation under the following
conditions:
Machine: 7.5 kW vertical mill
Workpiece material: 4130 steel, BHN 200
Cutting speed: 60 m/min
Feed: 0.25 mm per tooth
Depth cut: 5 mm
Cutting tool: 75 mm diam face mill, carbide teeth, 8 teeth.
7. Determine the required nose radius to obtain a surface finish of 100 mm rms
when turning AISI 4130 steel, 183 BHN, at 300 fpm and 0.20 feed, using
carbide cutting-tool material.
8. Determine the required nose radius to obtain a surface finish of 125 mm rms
when turning malleable cast iron, 120 BHN, at 45 m/min and 0.4 mm/rev feed,
using a high-speed-steel cutting tool.
9. What surface finish could be expected when turning AISI 1120 steel, 160
BHN, at 45 m/min and 0.25 mm/rev, using a high-speed-steel cutting tool with
1.6 mm nose radius.
382 Tool Design

10. Sketch a high-speed-steel single-point cutting tool with the following tool signa-
ture (use 10 mm by 10 mm shank): 0-8-24-8-10-0-1.6.
11. Sketch a brazed-type carbide tool with the following tool signature. Use a 25 by
30 mm by 200 mm tool shank. Use carbide-tool catalog to determine carbide
blank size. 0-6-7-7-15-15-0. 8.
12. A single-out cutting tool has a back rake of –5°, a side rake of 10°, and a
side cutting-edge angle of 15°. Solve for the true rake and angle of inclination
mathematically.
13. Check your answer to Prob. 12 by the nomograph in Fig. 4.66.
14. A high-speed-steel cutting tool has a back-rake of 16°, a side rake of 10°, and
a side cutting-edge angle of 30°. Solve for the true rake and angle of inclination
mathematically.
15. Check your answers to Prob. 14 by the nomograph in Fig. 4.66.
16. Determine the true rake and angle of inclination for a face-milling cutter with
positive axial of +15°, negative radial rake angle of –20°, and a 45° corner
angle.
17. A carbide face-milling cutter is listed in a catalog as having +5° positive axial
rake and –7° negative radial rake with 10 mm by 45° chamfer angle. (a) What
is the true rake and angle of inclination? (b) What cutting efficiency can be
expected?
18. A workpiece made of a tough alloy steel was to be machined with a carbide
face-milling cutter. A standard cutter with –7° radial rake and –7° axial rake was
selected to increase the strength of the cutting edge. The face-milling cutter
was available with a 15, 30 or 45° chamfer angle. The 15° chamfer angle was
agreed upon, and a cutter was purchased. When applied to the workpiece, the
15° chamfer face mill produced a finished surface that was burnished in places
because chips were apparently not clearing the cutter. Investigate the milling-
cutter geometry and make recommendations.
19. Determine the horsepower required and the thrust developed when drilling
grey cast iron with a 20 mm standard drill at a feed of 0.3 mm/rev and a speed
of 500 rpm.
20. A drill press is to be purchased for the toolroom. What horsepower rating should
be specified when the maximum size hole to be drilled will be 25 mm in tool
steel?
21. A C-1 grade of tungsten carbide is being used to machine cast iron. After being
in service for a short period of time, the cutting edge shows excessive flank
wear. Make recommendations.
22. A highly alloyed cast iron is being machined with a C-3 grade of tungsten car-
bide. A crater develops in the top of the insert when the chip flows over it after
a short period of time. Make recommendations.

Design Problems
1. Design a stub boring bar to machine a hole in a cast-iron gear housing. The
bore is 100 mm in diameter by 50 mm deep and is a through hole. The machine
tool is a horizontal boring mill.
2. A 40 mm-diam by 75 mm deep blind hole is being bored on a turret lathe
in 4130 steel. A 25 mm. boring bar made of 1040 heat-treated steel with a
Design of Cutting Tools 383

carbide cutting tool is used to bore the hole. The overhang of the bar is 90
mm. A considerable amount of chatter is present unless the cutting speed is
reduced to a figure so low that production is drastically reduced. Design a bor-
ing bar that will not chatter at production cutting speeds.
3. Design a nozzle system for a flood application of a cutting fluid for the gang-mill
setup shown in Fig. 4.150a. Two workpieces are machined at the same time.
4. Design a nozzle system for a flood application of a cutting fluid for the face-mill
setup shown in Fig. 4.150b.
5. Design a nozzle system for a flood application of a cutting fluid for the surface-
grinding set-up shown in Fig. 4.150c.
6. A single-point brazed carbide tool with a 15° side cutting-edge angle is being
used to machine AISI 1020 hot-rolled steel. At carbide cutting speeds the chip
is coming off the machine in a continuous blue-hot ribbon, causing a con-
siderable reduction in production because the operator is continually fighting
the chip or stopping the machine to remove a chip entanglement. Design a
new brazed-type carbide tool with provision for chip control. Shank dimensions
should be 12.5 by 12.5 mm.
7. A single-point brazed carbide tool with 0° side cutting-edge angle is being used
to machine AISI 1132 steel in an automatic lathe. A step-type chip breaker is
utilised to break the chip, but small fragments of the chip occasionally interfere
with the lathe mechanism and result in a scrapped part. Redesign the tool and
chip breaker to eliminate the small chip fragments. Tool-shank dimensions
should be 12.5 by 12.5 mm.

(a) (b)

(c)

Fig. 4.150 Design cutting-fluid nozzles for the above machines. (a) Two workpieces ma-
chined at the same time (b) 150 mm. carbide face mill (c) Surface grinder, 400
by 50 mm. grinding wheel
384 Tool Design

8. Design a single-point high-speed-steel tool to be used in a standard stub bor-


ing bar 12.5 mm in diameter. The bar contains a 6 mm. square hole at right
angles to the axis of the bar to hole a ingle-point tool. The single-point tool is
held in the 6 mm square hole with a setscrew. The boring bar is to be used
to bore a 22 mm through hole in aluminium. Design the tool to be set above
centreline with the proper geometry for on-centre cutting action. Determine the
geometry from Table 4.17.
9. Design a line boring bar to machine the bore of the cast-iron pulley housing
shown in Fig. 4.151. Both bores are to be machined in one pass. Design the
bar to be used in a horizontal milling machine. Use an existing machine in
the school shop or obtain milling-machine specifications from a catalogue or
brochure.

50
63

190

250

Fig. 4.151 Gray cast-iron pulley housing

10. Design a boring head or boring bar to machine the bore of the shaft carrier
shown in Fig. 4.152. Four shaft carriers are to be machined in one pass. Design
the boring head or bar to be used on a horizontal milling machine. Use an exist-
ing machine in the school shop or obtain milling-machine specifications from a
catalogue or brochure.
11. Design a line boring bar to machine the bores of the cast-iron gear housing
shown in Fig. 4.153. Utilise standard commercial block-boring tools or boring
cartridges in the design. The boring bar is to be used on a large horizontal mill-
ing machine or a horizontal boring mill. Use an existing machine in the school
shop or obtain machine specifications from a catalog or brochure.
12. Design a stub boring bar and holder for an engine lathe that can be used on
bores from 13 mm in diameter. Design the holder to mount on the compound
rest of the engine lathe and provide a method of adjusting the height of the bar
to coincide with the axis of the lathe. Use an existing lathe in the school shop
or obtain lathe specifications from a catalogue or brochure.
Design of Cutting Tools 385

150

Fig. 4.152 Shaft carrier flame-cut from 7.5 mm 1020 hot-rolled plate

460

40 Typ. 45 Typ.

305

230

380

20

Fig. 4.153 Gray cast-iron gear housing

13. Design a boring-bar holder for holding the preformed bars shown in Fig. 4.89.
The holder should hold bars from 6 mm to 20 mm in diameter and should
have provisions for adjusting the height and radial position because the correct
386 Tool Design

setting of this type of bar is quite important. Use an existing lathe in the school
shop or obtain lathe specifications from a catalog or brochure.
14. Design a reamer holder with angular and parallel float to be used on a turret
lathe. Design holder to accommodate taper-shank reamers up to a no. 3 Morse
taper. Use existing lathe in school shop for specifications or obtain lathe speci-
fications from catalog or brochure.
15. Your instructor will assign one of the workpieces shown in Figs 4.154 to 4.158.
Make a turret-lathe operation sheet for the assigned workpiece, selecting stan-
dard tools. Make a drawing of the tools and turrets similar to Fig. 4.159. Use
the notation shown in these illustrations and follow the same general methods.
Fine drawing technique required.

Neck 3 wide 35° chamfer to


1 R. Recess depth of thread
1.6 deep

30 20

6 Drill through
16 17
9.5 Drill, 25 deep
46

Fig. 4.154 Adaptor

6 × 45Chamfer 3.2 × 45 Chamfer

f 203 50 70 f 178 f 190

1
2

25
45

59
84

156

Fig. 4.155 Gland


Design of Cutting Tools 387

Fig. 4.156 Bolt

1.6 × 45° 1.6 × 45° Chamfer


Chamfer
10 Drill, 20 deep
12.70
12.60

19.10
25.4
25.3

19.00

15.90
15.80
33.40
33.10
32.50
52.40

Fig. 4.157 Pin


R3

22 41 102 127
0.8 × 45°
Chamfer 20
6

16
25

31

Fig. 4.158 Pulley


388 Tool Design

Fig. 4.159 Typical bar setup on turret lathe (Warner Swasey Co.)

References
Books
• ASTME: “Manufacturing Planning and Estimating Handbook,” F W Wilson
(ed.) McGraw-Hill, New York, 1963.
—: “Tool Engineers Handbook,” 2d ed., F W Wilson (ed.), McGraw-Hill, New
York, 1959.
—: “Machining with Carbides and Oxides,: Frank W Wilson (ed.), New York,
1962.
• Bhattacharyya, A, and I Ham: “Design of Cutting Tools,” SME, Dearborn, Mich.
1969.
• Brierley, R G, and H J Siekmann: “Machining Principles and Cost Control,”
McGraw-Hill, New York, 1964.
• Cincinnati Milling Machine Company: “A Treastise on Milling and Milling
Machines,” 3d Ed., Cincinnati, Ohio, 1951.
• LeGrand, R (ed.): “The New American Machinists’ Handbook,” McGraw-Hill,
New York, 1955.
• MCTI: “Metal Cutting Tool Handbook,” New York, 1965.
Design of Cutting Tools 389

• Oberg, E, and F D Jones: “Machinery’s Handbook,” 17th ed., The Industrial


Press, New York, 1965.
• SME: “Cutting and Grinding Fluid: Selection and Application,” Dearborn, Mich.
1967.
—: “Gundrilling, Trepanning, and Deep Hole Machining,” rev. ed., Dearborn,
Mich., 1967.
—: “Cutting Tool Materials Selection,: Dearborn, Mich., 1968.
—: “Fundamentals of Tool Design,” Prentice-Hall, Englewood Cliffs, NJ,
1962.
Catalogues
• Acme Machine Products Division, Muncie, Ind.
• American Broach & Machine Co., Ann Arbor, Mich.
• Armstrong Bros. Tool Co., Chicago.
• Barber-Colman Company, Inc., Rockford, Ill.
• Black Drill Co., Inc., Cleveland, Ohio.
• Carboloy Department, General Electric Company, Detroit.
• Cleveland Twist Drill Co., Cleveland, Ohio.
• Colonial Broach Co., Detroit.
• Davis Boring Tool Division, Giddings & Lewis Machine Tool Co., Fond du Lac,
Wis.
• Ex-cell-o Corp., Detroit.
• Geometric Tool Co., New Haven, Conn.
• Gisholt Machine Co., Madison, Wis.
• Greenfield Tap & Die Corp., Greenfield. Mass.
• Illinois Tool Works, Chicago.
• Kennametal, Inc. Latrobe, Pa.
• La Pointe Tool Co., Inc., Hudson, Mass.
• Lovejoy Tool Co., Springfield, Vt.
• McCrosky Tool Corp., Meadville, Pa.
• Modern Tool Works, Rockester, NY
• Morse Twist Drill & Machine Co., New Bedford, Mass.
• National Twist Drill & Tool Co., Detroit.
• O. K. Tool Co., Inc., Milford, NH
• Q-C Engineering Products, Detroit.
• Ready Tool Co., Bridgeport, Conn.
• Russell, Burdsall & Ward Bolt and Nut Co., Port Chester, NY
• Scully-Jones & Co., Chicago.
• Sheldrick Mfg. Co., Upper Sandusky, Ohio.
• Taft-Pierce Mfg. Co., Woonsocket, RI
• Union Twist Drill Co., Athol, Mass.
• Warner & Swasey Co., Cleveland, Ohio.
• Weldon Tool Co., Cleveland, Ohio.
5 GAUGES AND GAUGE
DESIGN

5.1 INTRODUCTION
Modern manufacturing requires extensive use of gauges for shop work, inspection
and reference. Shop gauges are used by workmen. Inspection gauges are used by
inspectors to check the manufactured product, and reference gauges are reserved for
checking the other two types.
A gauge is defined by the Sheffield Corporation as “a device for investigating the
dimensional fitness of part for a specified function.” Gauging is defined by the ANSI
as “a process of measuring manufactured materials to assure the specified uniformity
of size and contour required by industries.”
Basically, gauging accomplishes two things: (1) It controls the dimensions of a product
within the prescribed limitations, and (2) It segregates or rejects products that are
outside these limitations.
Gauging devices and gauging methods, like other phases of tooling in modern manu-
facturing, have become standardised. Generally speaking, standardised components
that can be obtained commercially are assembled into a unit to gauge a particular
product. It is therefore quite important that the tool designer be familiar with gauging
equipment and practise.
It may be necessary to design special gauges for checking dimensions that do not
readily adapt to standard gauges. A gauge of this type may be quite simple, as shown
in Fig. 5.1. Frequently, time can be saved by the use of a simple length gauge in
place of a machinist’s rule when a quantity of workpieces is involved. It should not be
assumed that special gauges are necessarily elaborate or that they are used only to
measure close tolerances.

Tolerance limits Gauge

Workpiece

+ 0.001
9.734
– 0.000
+ 0.000
9.765
– 0.000

Fig. 5.1 Simple length gauge

There are many gauging methods used to determine when a product conforms dimen-
sionally with drawings, specifications, or other prescribed requirements. However,
Gauges and Gauge Design 391

when these methods are analysed, it will be found that they are designed to check
one of the seven basic elements of workpiece geometry:

Distance Used to specify the relative location of the various components and ele-
ments of the workpiece. Distance is measured by comparison to a known standard.

Flatness Used to ensure that every element of a surface is within a specified dis-
tance from a nominal surface plane. Determines straightness and alignment of a
product.

Parallelism Used to ensure that two flat surfaces are parallel to each other.

Perpendicularity (squareness) Used to determine that two flat surfaces are normal
to each other.

Angularity Used to specify the angle between two flat surface, other than 90°.

Concentricity Used to ensure that points on a cylindrical surface are concentric to


a common axis.

Surface texture Similar to flatness but concerned with irregularities in a surface


rather than straightness and alignment.
These elements of workpiece geometry are theoretical and are unattainable in actual
machining practise. It is therefore necessary to specify the degree of variation (toler-
ance) that is acceptable for the product to function properly. Gauges are the means
by which the products are checked to determine whether the elements of geometry
fall within the specified variation.
There are many types of gauges, but they can all be placed in three broad groups:
(1) Fixed gauges, (2) Indicating gauges, and (3) Combination gauges. Fixed gauges
are the most common and are used for both large and small production. Fixed limit
gauges are used to ensure that a product is within the prescribed limits of size. Most
of these are standard gauges built to specifications of the Precision Gauge Design
Committee and manufactured in large quantities by companies specialising in this
type of tool. However, the tool designer is frequently given the job of designing simple
fixed gauges for special work.
Indicating gauges are comparatively complex. They measure the extent of deviation
from a normal or basic size and provide a means of evaluating the size of each part,
establishing its relationship to limits of size, and determining to what extent limits of
size are exceeded by rejected parts. In general, indicating gauges are placed in ser-
vice by the inspection department after consultation with gauge manufacturers.
Combination gauges are special devices designed to measure or check more than
one dimension of a workpiece at a given setup.

5.2 FIXED GAUGES


Fixed gauges may be classified as fixed master gauges or fixed-limit gauges. The
master gauge is made to represent the workpiece dimension in its nominal condi-
tion and is used as a setting gauge for setting up comparator-type measuring instru-
ments or as a reference standard for calibrating direct-measuring tools which require
392 Tool Design

periodic readjustment. Master gauges are dimensioned to represent the dimension to


be gauged. The dimension may be the basic size or the median size of the tolerance
zone.
Fixed limit gauges are used to determine whether the product is within prescribed
limits and are intended for use as inspection gauges. Since there is a high and low
limit on the product, two gauges are usually required, which are made to constitute
the design limits of that dimension.
Master gauges are usually designed and manufactured by companies specialising in
this type of product. The tool designer has little to do with master gauges. The follow-
ing discussion will generally be concerned with fixed limit gauges, since this type is of
more concern to the tool designer.
Fixed limit gauges have both advantages and disadvantages compared to indicat-
ing gauges. The use of fixed limit gauges requires greater sensitivity and skill on the
part of the inspector as there is an element of “feel” involved. They also fail to give
the amount of deviation from a nominal dimension in positive values, as indicating
gauges do. However, fixed gauges are essentially free from errors caused by drift
from original adjustment, backlash in mechanical movements, nonlinear response,
power fluctuations, and other factors which necessitate regular calibration of common
types of indicating gauges. Fixed limit gauges also give a positive yes or negative
no decision regarding the acceptability of the inspected part. They may be used in
any part of the manufacturing facility as they do not require a power supply. They are
quite rugged and remain accurate with reasonable care. They are less susceptible to
inaccuracy caused by dirty conditions and are relatively inexpensive to manufacture
or purchase.

Ring gauges These are gauges whose inside measuring surfaces are circular in
form, as shown in Fig. 5.2. Ring gauges are used to measure cylindrical surfaces,
tapers on shafts, and similar workpieces, as well as external threads. Large ring
gauges are turned down on the ends, as shown, to lighten them.

Receiving gauges These are similar to ring gauges but are used to verify the speci-
fied uniformity of size and contour of noncircular holes. They are extensively used to
check splined shafts. Three receiving gauges are shown in Fig. 5.3.

Fig. 5.2 Ring gauges (The Southern Gage Company)


Gauges and Gauge Design 393

Fig. 5.3 Receiving gauges

Plug gauges These are used to check the uniformity of holes. A plug gauge may
be straight or tapered and of any cross section. It may have either an integral or
replaceable handle. There are two types of replaceable handles: (1) The taper lock
(used for gauges up to 38 mm in diameter) and (2) The trilock (used for gauges above
38 mm in diameter). The former has a gauge member made with at taper shank that
fits tightly into a mating taper in the handle, while the latter has a reversibly gauging
member which has a central hole and three adjacent grooves. A screw passes through
the gauge and fastens this to the handle while at the same time ensuring a snug fit
by forcing three wedge-shaped prongs into the corresponding grooves of the gauge
(see Fig. 5.4). A drawing of a taper-lock plug gauge is shown in Fig. 5.5. Note that all
information necessary to make the tool is placed on the drawing.

ern
South
ern
South

hern
Sout

ern
South
Southern
ern
South
Southern

Fig. 5.4 Plug gauges (The Southern Gage Company)

Pin gauges When the holes are larger than 75 mm (as, for example, automobile
cylinder bores), plug gauges are very heavy. In these cases, it is more convenient
to use pin gauges. In using a pin gauge, the gauge is placed lengthwise across the
cylinder bore, and the measurement is made in a manner similar to that of an inside
micrometer. Pin gauges are also used to measure the width of slots or grooves. In
this connection, they are sometimes called width gauges. Figure 5.5 shows a typical
pin gauge.

Snap gauges These are fixed gauges with inside measuring surfaces for checking
diameters, lengths, thicknesses, and similar dimensions (see Figs. 5.6 and 5.7).

Length gauges A length gauge is a special gauge designed to replace a common


machinist’s scale in order to check a workpiece for length quickly as shown in
Fig. 5.1.
394 Tool Design

Drive and peen


Workpiece

Go Not go

Harden temper and grind

Fig. 5.5 Pin gauge

Fig. 5.6 Model A and B snap gauges (The Southern Gauge Company)

Fig. 5.7 Model C snap gauges (The Southern Gauge Company)

Flush-pin gauges These are used for gauging special shapes that may be difficult
to check by other means. They are extensively used for gauging the depth of slots.
Flush-pin gauges are not listed as standard in manufactures’ catalogues but are gen-
erally made for specific jobs by toolmakers. Figure 5.8 shows a typical example.
A definition given by the Bureau of Standards is
A flush-pin gauge is a gauge for checking the distance between two surfaces, com-
prising a body having a through hole, and a pin in the hole which projects from a face
of the body a distance equal to the dimension to be gauged when the opposite or
indicating end of the pin is flush with the opposite face of the body. The indicating end
of the pin, or the adjacent face of the body, has a step of a depth equal to the tolerance
on the dimension gauged.
Gauges and Gauge Design 395

+ 0.000
Max. 10.810
– 0.001
+ 0.001
Body Min. 10.790 Workpiece
– 0.000

Pin

Pin retaining screw


Pin to be flush with steps
for max. and min. positions

Enlarged view of body

Fig. 5.8 Flush-pin gauge

Flush-pin gauges are also called feeler-pin gauges. This type of gauge is particularly
useful for gauging the distance between surfaces that are not opposite to each
other.
Figure 5.10 shows a flush-pin gauge
designed to check the position of a
groom as turned workpiece. Figure 5-11
shows a drawing of a flush-pin gauge
used to mean the depth of drilled holes.
The general dimensions for this type of
gauge are
A = max. depth of component hole
A1 = min. depth of hole
B = min. diameter of hole
C = gauge max depth to apex
B
= __ tan f + A Fig. 5.9 Plug and ring thread gauges
2
(The Van Keuren Company)
C1 = gauge min. depth to apex

B1
= ___ tan f + A1
2
D = diameter of gauge pin = (B1 –0.002) – 0.001 tolerance
E = height of step = C – C1
Flush-pin gauges of this type should not be used for gauging diameter as tolerances
less than 0.125 mm except in special cases.
396 Tool Design

Thread gauges In 1920, Congress appointed the National Screw Thread Commission
to improve the practise of making and gaging screw threads. This led to the adop-
tion of two series of threads for the United States, the National Coarse (NC) and the
National Fine (NF).
The student should be familiar with the thread profile known as the American National
Form (ANF), which applies to all standard threads in the United States. The thread
profile is a modification of the V thread. It has an included angle of 60° and is flattened
at the crest and root to an amount equal to one-eighth of the pitch.
Five elements determine the fit of a threaded part with its mate and thus provide
the freedom or tightness required for correct performance: (1) Lead, (2) Angle, (3)
Roundness, (4) Thread form and (5) Diameter. The thread gauge is used to ascertain
whether these elements are acceptable.
In practise, fixed-thread gauges are in the form of ring or plug gauges. Figure 5.9
shows a thread plug gauge and a thread ring gauge. This type of gauge has a definite
limitation for very precise measurements. A single full-form plug or ring screw gauge
cannot check all the thread elements, as demonstrated in Fig. 5.10, where the thread
of a bolt is being checked in a ring thread gauge. This bolt would be passed by an
inspector relying on the feel of the fit, since the end threads are in good contact.
However, the bolt should be rejected because the pitch is incorrect. At least three
sets of plug and ring gauges are required to check a screw thread completely. One
set will check the pitch diameter, the second the major diameter, and the third the
minor diameter. In the case of an internal thread, a plain plug gauge would be used
to measure the minor diameter.

Nut-short lead
Bearing
Bearing

Screw

Bearing Nut-long lead Bearing

Screw
correct lead

Fig. 5.10 Checking a screw thread

In practise, ring and plug screw gauges are used for checking ordinary work, but
where great accuracy is required, other methods, such as the contour-projection
method, are employed.
Gauges and Gauge Design 397

Side clearance
angle

Front
clearance
angle
End cutting
edge angle Railroad wheel-tread and
flange contour

Fig. 5.11 Tool form gauge Fig. 5.12 Railroad wheel-tread


form gauge

Ways of machine
Shaft

Fig. 5.13 Shaft template gauge Fig. 5.14 Template gauge for machine ways

Form gauges These are specially designed to check the form or contour of a work-
piece. Consequently, this type of gauge is of particular interest to the tool-design
draftsman. Figure 5.11 shows a typical form gauge used to check a cutting tool.
Another form of gauge is the one shown in Fig. 5.12, used to check the contour of
railroad wheel treads. Pratt & Whitney makes a series of similar gauges to the speci-
fications of the American Association of Railroads. These gauges not only check the
contour of a tire or flange when new but show the amount of wear and indicate when
wheels should be reconditioned.

Template gauges These are similar to form gauges but are designed to check the
position and dimensions of two adjacent surfaces. Figure 5.13 shows a template

Go
69.84
69.85
69.39
69.40

Fig. 5.15 Go-not-go gauge Fig. 5.16 Adjustment of snap gauge (Pratt
& Whitney Corp.)
398 Tool Design

gauge designed to check two shoulders of a shaft. Figure 5.14 shows another tem-
plate gauge designed to check the ways (or slides) of a machine tool.

Go-not-go gauges Twin gauges measuring the maximum and minimum limits of a
workpiece are called go-not-go gauges. A simple snap gauge of this type is shown
in Fig. 5.15. Here the upper limit is 69.85 mm less 0.01 mm tolerance and the lower
limit is 69.39 mm plus 0.01 mm tolerance. This means that a workpiece 69.86 mm or
over will not enter the gauge and should be rejected. Also a workpiece under 69.39
mm will enter the second part of the gauge and should also be rejected. Thus, only
workpieces going into the first part but not going into the second part are accepted.
This type of snap gauge is known as a solid progressive snap gauge. Note that the
gauging surfaces cannot be adjusted for wear or to obtain different fits between the
workpieces being checked. The gauges shown in Fig. 5.16 are adjustable. It is the
usual practise to have responsible persons make the adjustments and then lock and
seal the screws before giving them to the inspectors for routine work. Figure 5.16
shows the adjustments of the Pratt & Whitney Trusform snap gauge. In this type, the
adjusting screw has right- and left-hand threads which work in combination to push or
pull the anvil 1.27 mm per turn. The locking device consists of two bevelled bushings
which clamp against the flat surfaces on the anvil shank and hold it solidly. The device
used on this adjustment to obtain a very fine setting is called a differential screw.

Fig. 5.17 Dovetail gauge


Gauges and Gauge Design 399

Dovetail gauges Dovetail gauges used to measure V slides are shown in Fig. 5.17.
Since the ends of the V ’s are relieved, it is impossible to ensure two mating surfaces
by checking the dimensions from edge to edge. Consequently, two rollers or buttons
are placed in the V of the gauge, and the dimensions X and M are measured as
shown. If these and the angle are correct, then the gauging surfaces are also correct.
The calculations for the measurements X and M are shown below. From Fig. 5.17a,
D
C = __ cot A
2
where D is the diameter of roller
C = 7.626 cot 27.5 = 7.626 × 1.9210 = 14.6495 mm
X = D + 2C + 116.77 = 15.252 + 29.299 + 116.77 = 161.321 mm

Fig. 5.18 Sine-bar vise (Universal Vise


and Tool Company)

25.47
L = 2 × ______ + 116.77
tan 55°
= 35.6685 + 116.77 = 152.4385 mm
M = 151.38 – (2C + 15.252) Fig. 5.19 Checking a taper (CEJ
= 151.38 – (2 × 14.6495 + 15.252) Gage Company)

= 106.829 mm

Taper gauges Tapered holes are commonly checked by tapered plug gauges.
Tapered shafts and tapered plug gauges can be checked by means of a sine bar.
Figure 5.18 shows a commercial since bar. Figure 5.19 shows how the measurements
are made.

Gauge blocks Length dimensions used in industry are based on the standard unit of
dimensional measurement, such as the international inch or meter. Until recent times,
the basic reference unit has been the International meter bar or its official duplicates.
More recently, the international basic standard of length has been defined in terms
of a specified wavelength of light. In theory, all gauges should be checked against
the basic international standards, but when millions of gauges are considered, such
a procedure would obviously be impractical. Large companies may be able to stand
400 Tool Design

the expense of special equipment needed for achieving reliable length measurements
based on wave-length, but the majority of companies cannot. The logical solution is
to use the basic standard to calibrate many secondary standards, which are close
replicas of the former. The secondary standard would be manufactured in quantity by
mass production methods in order to be relatively inexpensive.
The secondary standards that have emerged are known as gauge blocks and have
become a commonplace in modern industrial plants. On first thought, one tends to
classify gauge blocks as a master fixed gauge, but in reality they are much more than
this. A single gauge block may be classed as a master fixed gauge, but a single gauge
block has little use except in special cases. Gauge blocks must be viewed from the
concept of several single master gauge blocks being combined into a single gauge
bar. The combined single blocks result in a bar whose actual dimension truly repre-
sents, within specific limits, the nominal dimension sought for a particular application.
A typical set of gauge blocks is shown in Fig. 5.20.

Fig. 5.20 Typical sets of gauge blocks (Pratt & Whitney, Inc.)

The tool designer will not be responsible for the manufacture of gauge blocks, but
needs to understand how they are used. The gauges he designs will be calibrated
and checked with gauge blocks before the gauge is placed in use.
Gauge blocks or slip gauges are made of steel, chrome-plated steel, stainless steel,
tungsten carbide, or chrome carbide, depending upon the manufacturer and the price
the purchaser is willing to pay. They are stabilised for minimum dimensional change
and are hardened to RC63 to 64. Standard blocks may be rectangular or square in
shape. The gauging surfaces are lapped to a very high finish.
Gauge blocks are manufactured worldwide using various standards. Gauge block
sets are available with various numbers and combinations of blocks. There are two
basic ways of designating gauge block sets; inch and mm units or metric units. The
inch unit gauge block sets are designated by letter E. Some of the commonly used
inch unit gauge-block sets are E28, E35, E41, E49, and E81. A standard E81 block
sizes are shown in Table 5.1. The metric unit gauge-block sets are designated by let-
ter M. Some of the commonly used metric unit gauge block sets are M27, M33, M50,
M87, M105 and M112. A standard M112 block sizes are shown in Table 5.2.
In order to select the gauge blocks required to produce a certain dimension, subtract
or eliminate the last digit of the length. The objective is to make up the stack of as few
as possible. The procedure is illustrated in Example 5.1.
Gauges and Gauge Design 401

Table 5.1 Arrangement of gauge block sizes in a standard E83-block set


First series: ten-thousandths-9 blocks
0.1001 0.1002 0.1003 0.1004 0.1005 0.1006 0.1007 0.1008 0.1009
Second series: thousandths–49 blocks
0.101 0.102 0.103 0.104 0.105 0.106 0.107 0.108 0.109 0.110
0.111 0.112 0.113 0.114 0.115 0.116 0.117 0.118 0.119 0.120
0.121 0.122 0.123 0.124 0.125 0.126 0.127 0.128 0.129 0.130
0.131 0.132 0.133 0.134 0.135 0.136 0.137 0.138 0.139 0.140
0.141 0.142 0.143 0.144 0.145 0.146 0.147 0.148 0.149
Third series: fifty-thousandths–19 blocks
0.050 0.100 0.150 0.200 0.250 0.300 0.350 0.400 0.450 0.500
0.550 0.600 0.650 0.700 0.750 0.800 0.850 0.900 0.950
Fourth series inches–4 blocks
1.000 2.000 3.000 4.000

Table 5.2 Arrangement of gauge block sizes in a standard M112 block set
First Series: 9 blocks
1.001 1.002 1.003 1.004 1.005 1.006 1.007 1.008 1.009
Second Series: 49 blocks
1.01 1.02 1.03 1.04 1.05 1.06 1.07 1.08 1.09
1.1 1.11 1.12 1.13 1.14 1.15 1.16 1.17 1.18
1.19 1.2 1.21 1.22 1.23 1.24 1.25 1.26 1.27
1.28 1.29 1.3 1.31 1.32 1.33 1.34 1.35 1.36
1.37 1.38 1.39 1.4 1.41 1.42 1.43 1.44 1.45
1.46 1.47 1.48 1.49
Third Series: 49 blocks
0.5 1 1.5 2 2.5 3 3.5 4 4.5
5 5.5 6 6.5 7 7.5 8 8.5 9
9.5 10 10.5 11 11.5 12 12.5 13 13.5
14 14.5 15 15.5 16 16.5 17 17.5 18
18.5 19 19.5 20 20.5 21 21.5 22 22.5
23 23.5 24 24.5
Fourth Series: 4 block
25 50 75 100
Fifth Series: 1 blocks
1.0005

You are required to produce a dimension of 53.717 mm using


Example 5.1 M112 gauge block set. Select the minimum number of gauge
blocks to produce the dimensions.
Solution: Desired dimension 53.717
Select block to eliminate the 0.007 mm digit – 1.007 Block 1
52.710
402 Tool Design

Select block to eliminate the 0.01 mm digit – 1.010 Block 2


51.700
Select block to eliminate the 0.7 mm digit – 1.700 Block 3
50.000
Eliminate 50 mm – 50.000 Block 4
0.000
This gives the combination of four blocks to measure the dimension 53.717 mm.
However, other combinations are also possible.
A 100 mm sine bar is to be set to an angle of 45° 45¢ 18≤. Find
Example 5.2 the higher of the gauge blocks required. Also select the mini-
mum number of gauge blocks to produce the height using M112
gauge block set.
Solution: For a sine bar the sine of the angle is equal to the ratio of the height by the
sine bar length. That is,
H
sin q = __
L
Given q = 45° 45¢ 18≤ = 45.755°, L = 100 mm
Therefore,
H = L sin q = 100 × sin 45.755 = 71.636 mm
The height of the gauge block required is 71.636 mm.
The minimum number of gauge blocks to produce the height using M112 gauge block
set, can be found out as follows:
Desired dimension 71.636
Select block to eliminate the 0.006 mm digit – 1.006 Block 1
70.630
Select block to eliminate the 0.03 mm digit – 1.030 Block 2
69.600
Select block to eliminate the 0.6 mm digit – 1.600 Block 3
68.000
Eliminate 18 mm – 18.000 Block 4
50.000
Eliminate 50 mm – 50.000 Block 5
0.000
In order to achieve the height of 71.636 mm, 5 blocks (i.e., 1.006, 1.03, 1.6, 18,
50 mm) are required to be stacked.
Gauges and Gauge Design 403

When individual gauge blocks are combined, or built up, to provide a specific mea-
surement, they must be wrung together. This is accomplished by a twisting and slid-
ing motion that squeezes out the air between the almost geometrically flat gauging
surfaces. When properly wrung together, the blocks adhere to each other to the extent
that considerable force must be exerted to pull them directly apart. This phenome-
non is usually explained as a combination of molecular attraction and the cementing
action of oil or moisture film on the gauging surfaces. Steel gauge blocks should
not be wrung together any longer than necessary, as moisture trapped between the
blocks may cause corrosion.
Gauge block sets are classed in grades according to accuracy. Grades are denoted
as Calibration Grade, Grade-00, Grade-0, Grade-I, Grade-II. Calibration Grade and
Grade-00 are used along with high magnification comparators. Grade-0 blocks are
used for inspection and accuracy works. Grade-I blocks are used for measurements
of tools and components. Grade-II blocks are used for wide tolerances work.

Optical flats Gauge blocks and other fine surfaces can be tested for flatness with
an optical flat, a special block of glass with one side very accurately ground flat. The
other side is flat but not parallel to the first. Both sides are polished and clear. If the
accurate face is placed on the test surface, light will be reflected from the two surfaces
in contact. The light waves reflected from one surface may be in step or out of step
with those reflected from the other. If the two surfaces are not in close contact, the
reflected rays will interfere with each other, causing dark spots or bands to appear.
These bands are curved, and the greater the curvature, the greater the departure of
the test surface from a true flat. If the test surface is truly flat, the interference bands
will be sharply straight and parallel. A special monochromatic light (one colour, i.e.,
one wavelength) is necessary. Special lamps, complete with transformer, are manu-
factured for his purpose. Light-wave interference produced by the optical flat can also
be used to compare the size of an unknown block with that of a known one.

5.3 GAUGE TOLERANCES


A fixed gauge, like other manufactured products, must be made to a manufacturing
tolerance. In general practise, when no tolerance has been specified, gauge toler-
ance is held to 10 per cent of the product total tolerance. This is often referred to as
the rule of 10 to 1. In other words, the inspection gauge or instrument should be 10
times more accurate as the dimension being measured.
It is a commercial practise to apply unilateral tolerancing to go and not-go gauge
members. The tolerances on go plain plug gauges is applied plus. Not-go plain plug-
gauge tolerances are applied minus. The go plain ring tolerances are applied minus,
not-go plain ring tolerances are applied plus. There are exceptions to this, one being
master rings and plugs when bilateral tolerances are commonly used; a second is
where encroachment on product lines is minimised by applying tolerance in not-go
gauges bilaterally.
The direction of tolerance is shown in Fig. 5.21.

5.4 THE SELECTION OF MATERIAL FOR GAUGES


Many factors affect the selection of material for gauges. The tolerance to be checked,
the number of items to be gauged, the composition and hardness of the material to
404 Tool Design

Fig. 5.21 Methods of allocating gauge tolerance [From ASTME, “Tool Engineers Handbook,”
2d ed., F W Wilson (ed.) McGraw-Hill, New York, 1959, by permission]

be gauged, the cost of the gauge material, and the complexity of the workpiece to be
gauged are the most important.
The major properties of materials from which gauges are manufactured are dimen-
sional stability and wear resistance. Wear resistance of a gauge depends upon sur-
face hardness, composition of the material, and the quality of the surface finish. The
surface must be hard but not so brittle that it will chip or break away.
Materials used for gauge construction are given in Table 5.3. Carbon steels and alloy
steels should be subjected to a deep-freeze treatment in addition to heat treating
to increase the dimensional stability of the gauge. This is especially important for
the higher alloy steels that are known to retain a high percentage of austenite after
quenching. Austenite that is retained in a finished gauge will gradually transform dur-
ing service life, resulting in dimensional change of the gauge.
The amount of autenite transformed can be increased by subzero cooling (deep
freezing). This condition promotes stability because it reduces the amount of austen-
ite which later might transform and cause the finished gauge to grow. Subzero cooling
should take place immediately after the initial tempering that follows quenching from
the austenitising (hardening) temperature. Subzero temperatures should be from –32
to –49°C. Media for subzero quenching include freezer chests or low-temperature
liquids. A mixture of acetone and dry Ice will provide a temperature of 38°C.
Tungsten carbide materials for solid gauges provide very good stability and wear
resistance. However, their relatively high price and sensitivity to chipping limit their
Table 5.3 Recommended materials and hardness for most common types of gaugesa
Type of gauge Size of piece to be gauged. 1.27 Size of piece to be gauged,12.71 to Size of piece to be gauged, 12.71 to
to 12.71 mm, total tolerance, 0.013 mm 101.68 mm, total tolerance, 0.015 mm 101.68 mm total tolerance, 0.05 mm
Occasional Frequent All gauging Occasional Frequentc All gauging Occasional Frequent All gauging
gauging of gauging of of materials gauging of gauging of of materials gauging of gauging of of materials
metals softer metals softer harder than metals softer metals softer harder than metals softer metals softer harder than
than 350 BHN 350 BHN 350 BHN than 350 BHN than 350 BHN 350 BHN than 350 BHN than 350 BHN 350 BHN
Cylindrical, O2 or W1 at M2 at Salt-nitrided W1 at Salt-nitrided Tungsten W1 at Salt-nitrided Salt-nitrided
ring and plug RC61-64 RC61-64 M2 at carbideb RC6-64 M2 at M2 at
RC62-65 M2 at or carburised RC62-65 RC62-65 RC62-65
RC62-65 B1112c or D2
or tungsten
carbideb
Adjustable
snap:
Body Stress-relived cast iron, ASTM A48, class 25 to 30
Pins,
buttons and
anvils O2 at D2d Tungsten O2 at A2d Tungsten O2 at A2 at Salt-nitrided
RC61-64 carbideb RC61-64 carbideb RC61-64 RC62-65 M2 at
RC62-65
Height and
length O2 at M2 at M2 at W1 at A2d M2 at W1 at A2 at M2 at
RC61-64 RC62-65 RC62-65 RC61-64 RC62-65 RC61-64 RC62-65 RC62-65
Feeler O2 at L7e at D2 at
RC45-50 RC45-50 RC45-50
Pin gauge O2 or W1 at Salt nitrided Salt-nitrided O2 to W1 at D2d Tungsten O2 or W1 at A2 at Salt-nitrided
RC61-64 M2 at M2 at RC61-64 carbideb RC61-64 RC62-65 M2 at
RC62-65 RC62-65 RC62-65
Gauge blocks O2 at D2d D2d L7e at D2d Tungeten L7e at M2 at Salt-nitrided
RC61-64 Tungeten RC61-64 carbideb RC61-64 RC62-65 M2 at
carbideb RC62-65
Gauges and Gauge Design 405
Type of gauge Size of piece to be gauged. 1.27 Size of piece to be gauged,12.71 to Size of piece to be gauged, 12.71 to
to 12.71 mm, total tolerance, 0.013 mm 101.68 mm, total tolerance, 0.015 mm 101.68 mm total tolerance, 0.05 mm
Occasional Frequent All gauging Occasional Frequentc All gauging Occasional Frequent All gauging
gauging of gauging of of materials gauging of gauging of of materials gauging of gauging of of materials
metals softer metals softer harder than metals softer metals softer harder than metals softer metals softer harder than
406 Tool Design

than 350 BHN 350 BHN 350 BHN than 350 BHN than 350 BHN 350 BHN than 350 BHN than 350 BHN 350 BHN
Receiving, oval O2 or W1 at L7e at Salt-nitrided O2 or W1 at L7e at Salt-nitrided O2 at A2 at Salt-nitrided
and square RC61-64 RC61-64 M2 at RC 61-64 RC61-64 M2 at RC61 to 64 RC62-65 M2 at
RC62-65 RC62-65 RC62-65
Thread, ring O2 at A2d M2 at W1 at A2d M2f at W1 at A2 at M2f at
and plug RC62-64 RC62-65 RC61-64 or RC62-65 RC61-64 or RC62-65 RC62-65
carburised carburised
8620c 8620c
Roll thread,
snap:
Body Stress-relieved cast iron, ASTM A48, class 25 to 30
Rolls O2 at L7e at L7e at L7e at L7e at D2d L7e at L7e at D2 or M2f at
RC61-64 RC61-64 RC61-64 RC61-64 RC61-64 RC61-64 RC61-64 RC62-65
Spline, male
and female O2 at O6 at A2d O2 at O6 at A2d O2 at O6 at A2 or D2 at
RC61-64 RC61-64 RC61-64 RC61-64 RC61-64 RC61-64 RC62-65
Alignment … … … Carburised Carburised Carburised Carburised Carburised Carburised
bars B1112c B1112 or 8620c B1112c B1112 or 8620c or
8620c nitrided 8620c nitrided
4140g 4140g
a
Where more than one tool material is recommended for a specific set of conditions, the second listed is preferred for the larger sections of the indicated size range.
b
Group 2 carbide is usually recommended.
c
A case not more than one-fifth of the section thickness, equivalent to RC60 minimum is recommended.
d
For close tolerances, steel must be tempered to secondary hardness for maximum stability. Hardness values may be as low as RC56.
e
Tonnage steel 52100 has proved a satisfactory substitute for L7.
f
Salt nitriding is recommended for coarse pitches only.
g
Pretreat to R26 to 32 then gas-nitride for 25 hr.
SOURCE: From ASM, “Metals Handbook,” vol. 1, “Properties and Selection,” 1961, by permission.
Gauges and Gauge Design 407

use to application where extended use or abrasive conditions are prevalent. Because
the coefficient of thermal expansion of tungsten carbide differs from that of steel, a
controlled temperature environment may be required when gauging tightly toleranced
dimensions on parts made of steel. A change in room temperature may cause the
steel workpiece to expand or contract out of tolerance, whereas the carbide gauge
would not change the same amount.
Chromium plating gauges increases wear resistance and adds some resistance to
corrosion. It is also useful for restoring worn gauges to the original size in a salvaging
process. However, gauges with wear areas around sharp corners or edges should not
be chromium-plated because the plating may flake away.

5.5 INDICATING GAUGES


Indicating gauges detect variations in a specific distance and display them on a dial
or graduated scale. They have the ability to detect minute errors because the variation
shows on the scale in an amplified version.
The majority of indicating gauges compare the actual dimension of the workpiece with
the dimension of a master setting gauge. For this reason, indicating gauging methods
are often referred to as comparative gauging. The master setting gauge is equal to the
nominal dimension to be gauged. The indicating gauge actually measures only the
amount and direction of any deviation which exists in relation to the nominal size.
The amplification of indicating gauge movement may be mechanical, pneumatic, opti-
cal, or electric. The tool designer will seldom be required to design the amplification
system, as this is generally down by a company specialising in the manufacture of
indicating gauges. The tool designer’s job will be to design fixtures to adapt the par-
ticular amplification system to a practical inspection setup.

Mechanical indicators Mechanical indicators operate by different systems of dis-


placement amplification, which may consist of simple levers, cams, torsion strips,
gear trains, reeds, or a combination of these and other systems. Figure 5.22 shows
a simple level magnifier. The ratio of magnification is L2/L1. The accuracy is affected
by wear on the pivot at A, friction, and the inertia of the moving parts. The point B is
moved a small amount in the gauging process, and this is indicated on a scale at C in
terms that can be read quickly and easily.
Figure 5.23 shows a series of levers which is an extension of the simple lever move-
ment making a compact device with greater magnification. However, in this type the
errors due to wear, friction, and inertia must be carefully considered.

L2
L1

C
B A

Fig. 5.22 A simple lever magnifier


408 Tool Design

Fig. 5.23 A series lever magnifier

Figure 5.24 shows the lever movement applied to a simple depth gauge. The device
quickly checks the depth of surface A, and if the workpiece is resting on a flat surface
such as a surface plate, the uniformity of surface A can be checked by rotating the
workpiece. The spring ensures that the spindle will remain in contact with the surface,
and the pin located in the plunger slot prevents the plunger from falling out when the
workpiece is removed.

Fig. 5.24 Indicator depth gauge

The dial indicator is probably the most common


mechanical indicating gauge (see Fig. 5.25). Dial
indicators generally use some form of gears and
levers, as shown in Fig. 5.26. Various arrange-
ments of gears and levers are used, but basically
the amplification of the gauge is obtained solely
by the ratio of the various pinions, gears, and
levers, plus the length of the indicating pointer.
The dial-indicator gauge is rigidly mounted on
a gauge stand or in fixed position on a gauging
fixture. It shows dimensional deviations of the
Fig. 5.25 A typical dial indicator
part being gauged with respect to a reference
(Federal Products Corp.)
surface.
Gauges and Gauge Design 409

Dial indicators have the advantage of being easily mounted to gauging fixtures and
are completely independent of an external power source. They are not affected by
drift, which sometimes occurs in gauges whose operation is based on transducing the
linear displacement of the probe into electrical or fluid-pressure variations. They have
an extremely long gauge range relative to amplification. They are, however, somewhat
limited in accuracy when it is necessary to gauge extremely close variations.

Spring Case
Bearing

Gear
segment Pinion

Friction
spring Actuating
lever

Spring
Pivot
Collar

Steam

Anvil

Fig. 5.26 Dial-gauge mechanism

Dial indicators are frequently built into


gauging fixtures for gauging a par-
ticular workpiece. Figure 5.27 shows
a fast-operating indicator plug gauge
using a dial indicator.
Another good example of a gauging
fixture designed to use a dial indica-
tor as its magnification movement is
shown in Fig. 5.28. The workpiece is
an automobile piston. This is placed on
the V-locating surfaces G, and the bot-
tom of the piston-pin hole is pressed
firmly against the stationary locat-
ing pin L. The movable pin N is then Fig. 5.27 Indicator plug gauge used for the in-
allowed to touch the top of the piston- spection of holes (Federal Products
pin hole, as shown, and the position of Corp.)
410 Tool Design

pin N as indicated on the dial gauge checks the squareness of the hole with the walls
of the piston.

L
G

Fig. 5.28 Squareness-indicating gauge (Federal Products Corp.)

Figure 5.29 shows a simple fixture for checking the parallelism of rolls. The fixture can
be adjusted for a series of rolls to a maximum diameter of 24 in. The contact strips B
and C have notches to which the gauge is set for various sizes of rolls. When this has
been done, the indicator is moved downward by releasing screw I and thus freeing
bar D. The comparative diametrical readings are found by shifting the gauge to differ-
ent portions of the roll. For a truly parallel roll, the readings will be the same.

Fig. 5.29 Indicating parallel gauge (Federal Products Corp.)

The principle of the reed movement is shown in Fig. 5.30. The major part of this
device is an alloy-steel reed or flexible strip. Two means of magnification, mechanical
Gauges and Gauge Design 411

and optical, are combined to form a single unit.


Essentially, the device consists of two metal blocks,
one fixed and one floating, joined by the reeds. The
fixed block is rigidly anchored to the gauge-head
case. The floating block carrying the gauging spindle
is connected horizontally to the fixed block by two
reeds. A vertical reed is attached at the top of each
block to its inner side, and the upper ends of these
reeds are joined together. Beyond this joint extends
a pointer or target. The gauging spindle is an inte-
gral part of the movable block. When the spindle
and block are moved upward in a gauging operation,
the horizontal reed deflects slightly but the vertical
reed on the floating block tends to slip past its com-
panion. However, as these vertical reeds are joined
together at their upper end, instead of slipping the
movement causes both reeds to swing through an
are, and as the target is merely an extension of the
vertical reeds, it swings through a much wider are.
The amount of target swing is proportional to the Fig. 5.30 Principle of reed
distance the floating block is moved but of course magnifier
is very much greater. Through a series of lenses a
light beam projects the but of course is very much greater. Through a series of lenses
a light beam projects the shadow of the target on the scale of the gauge. Figure 5.31
shows a reed movement applied to a Sheffield visual gauge.
Figure 5.32 shows the principle of the twisted-strip mechanism. This type of mechani-
cal indicator employs 100 per cent mechanical amplification without gears or pivots.

B
A

Fig. 5.31 Visual gauge using reed mech- Fig. 5.32 Principle of twisted-strip
anism (Bendix Corporation, A mechanical indicator
& M Division.)
412 Tool Design

When the spindle A is displaced, it causes the spring knee B to move, which in turn
applies a tensile force along the axis of the twisted strip C. The resulting elongation
causes a change in the pitch of the twisted strip, which is transmitted to the pointer
D.
A very high degree of accuracy characterises this type of mechanism. There are
practically no wearing elements, and therefore the original accuracy can easily be
maintained. It is extremely sensitive and has a low gauging pressure because of
the almost frictionless operation of the twisted bands. It will not, however, measure
wide size variations.

Pneumatic gauges Pneumatic gauging is a means of measuring dimensions by air.


Basically, in pneumatic gauging, the nearness of a surface is detected by a jet of air
directed toward it by a nozzle which never touches it. The restrictive effect produced
as the surface approaches the nozzle has a well-defined relation to the clearance
between the two. This effect can be amplified and transmitted to a scale for measuring
purposes.
This principle is illustrated in Fig. 5.33. Here an air plug, usually containing two dia-
metrically opposed nozzles, is used to measure the inside diameter of a hole. This
type of air-gauging head is probably the most common and widely used in industry
today. The flow of air is metred through the air nozzles, and any change in the clear-
ance between the air nozzles and the hole walls will cause a change in the flow of
air. The air plug need not be accurately centred within the hole, for any decrease in
clearance to one nozzle will be compensated for by an increase in the clearance to
other if the nozzles are equally matched.
Referring again to Fig. 5.33, note that the air-escape orifices are recessed below
the cylindrical surface of the gauging head so that they never contact the part being
gauged. Gauging-head wear will not affect accuracy until the gauge head is worn
down to the orifice level.

Plug clearance Jet clearance Air escape

Gaging
head Airflow

Fig. 5.33 Basic principle of pneumatic gauging

The most common methods of detecting an change in flow of air that is metred through
the air nozzles is to actually measure the flow of air that is metred through the orifice
and to measure the variations of pressure in the gauging circuit. Figure 5.34 shows
the method of measuring the flow of air. Compressed air enters the gauge through
a filter and an automatic compensating pressure regulator. It then passes through a
Gauges and Gauge Design 413

tapered glass tube containing a small metal float, and then to the air nozzles con-
taining two diametrically opposed orifices for air escape. When the workpiece to be
gauged is placed over the gauging head (air nozzle), the flow of air is restricted and
changes the position of the float in the tube. The setting of the amplification is accom-
plished by two adjustments. One is the atomospheric bleed, which positions the float-
ing indicator, and the other is a magnification-adjusting screw for changing amplifica-
tions. It is necessary to have two master ring gauges for setting amplification. Figure
5.35 shows a typical flow-air gauge.

Fig. 5.34 Principle of flow-type air gauge

Figure 5.36 shows the principle of the pressure-type pneumatic gauge, which is often
called a back pressure gauge because the indicator deflects according to back-pres-
sure changes built up in the circuit when the workpiece is placed over the gauging

Regulator
Filter
Factory Indicator
air line Bourdon
tube
Mechanical
amplifying
device
Adjustable
restrictor
Metering
orifice
Fig. 5.35 Flow-type air gauge (The Fig. 5.36 Principle of back-pressure air gauge
Sheffield Corporation)
414 Tool Design

head. The deflection is amplified by a


lever-and-gear arrangement similar to
that in mechanical dial indicators.
In use, the air enters through a filter and
a regulator for controlling pressure. It
then passes a restriction for metreing
the air and then to the gauging head.
A Bourdon tube is placed in the circuit
behind the metreing restriction to indi-
cate pressure changes due to varying
the clearance between the orifice and
the part being gauged. The pressure
changes in the Bourdon tube causes it
to deflect. The deflections are amplified
by the gear-and-lever arrangement. The Fig. 5.37 Back-pressure air gauge (Federal
regulator which varies the air pressure Products Corporation)
and the restriction for metreing the air
are used to adjust amplification. Two master ring gauges are required to set amplifica-
tion. Figure 5.37 shows a typical pneumatic-gauge system utilising the back-pressure
principle.
Air-gauging methods are not limited to measuring inside diameters. It is just as easy
to measure outside diameters by placing two opposed air nozzles in a ring gauge.
Figure 5.38 shows typical air-gauging applications. It can readily be seen that the use
of pneumatic gauging is limited only by the ingenuity of the tool designer.
The major advantages of pneumatic gauges are as follows:
1. Simplicity: Does not have a great number of moving parts to wear. Production
workers do not require special training to use.
2. Deep bores or holes can be gauged.
3. Can be used without danger in an explosive atmosphere.
4. Does not mar surface finish or distort fragile workpieces.
5. Excellent for gauging rough surfaces.
6. Extremely fast and accurate.
7. Has a cleaning effect which reduces possibility of false readings due to oil, dirt,
or contact on the workpiece.

Electric and electronic indicators The basic principle of electric and electronic
gauges involves the displacement of a mechanical-contact type of sensing mecha-
nism. An electromechanical transducer converts the displacement to proportional
changes in electrical current, which are displayed on a metre calibrated in length
units. By proper circuitry, the indicating metre can be made to indicate plus or minus
directional movement from the centre in addition to magnitude.
The various types of electromechanical transducers used in electric and elec-
tronic gauges will not be discussed at any length. The electrical magnifier shown in
Fig. 5.39 will serve to give a general idea of the electromechanical transducer. When
the spindle is moved a small amount, the steel armature is moved in turn. The move-
ment of the armature generates an electric current in the coils according to the prin-
Gauges and Gauge Design 415

ciple of Lenz’s law. The voltage of this current is used to actuate a metre (calibrated
in length units) according to the magnitude of the voltage.

(a) (b) (c)

(d) (e) (f)

(g) (h) (i) (j)

(k) (l) (m)

Fig. 5.38 Typical manual pneumatic gauging operations. (a) ID, two-nozzle: measures ID,
bell mouth, out-of-round, taper. (b) ID, three-nozzle: measures average ID and
three-lobing. Also made for five- and three- or five-lobing. (c) OD, ring type.
Also made in snap type. Measures OD, out-of-round, and taper. (d) Chord types
measures OD only. Usually for large diameter. (e) OD, annular nozzle: measures
average OD of small parts. Also made for ID measurements. (f) Straightness:
measures straightness only. Independent of any diameter variations. (g) Taper, ID
or OD: measures taper, also large and small diameters. (h) Squareness: measures
squareness of face to bore by rotation. (i) Thickness: measures thickness only.
(j) Height, flatness: measures height, flatness, parallelism. (k) Matching: measures
clearance or interference of parts. (l) Centre distance: measures centre distance
only, independent of diameter. (m) Concentricity: measures concentricity as part
is rotated. (From American Machinist, October 1964, by permission)

The difference between electric and electronic gauges is in the degree of magnifica-
tion and the type of output. Both types rely on mechanical contact with the work to be
measured. The distinction between electric and electronic gauges is not well defined,
416 Tool Design

but generally the output of electronic gauge heads is amplified to a greater magnitude
before it is indicated on the metre scale.

Coil adjustment Steel armature


screw

Adjustable
coil

Coil adjustment
screw Flexible
strip
Stop screw
Locknut

Spindle

Fig. 5.39 An electric magnifier

The major advantage of electronic gauges is that they are quite sensitive and will
react to a very small mechanical input. Some instruments have the ability to sense
size differences and to amplify them. These characteristics are best put to use in a
laboratory measuring instruments such as gauge-block comparators and in compara-
tors and height gauges for highly precise shop applications. Figure 5.40 shows a typi-
cal electronic indicator.

Fig. 5.40 The Opt-O-Limit electronic indicator (Colt Industries, Pratt & Whitney Tool and
Gage Division)
Gauges and Gauge Design 417

Optical projectors Forming tools, gear teeth, fine threads, and similar parts can
be quickly inspected by means of optical projection. The term optical comparator or
projector is used for an instrument designed to magnify optical images of tools or
small parts. While the purpose of these instruments is similar to that of a microscope,
the optical projector handles much larger work than the microscope and in addition
throws an enlarged image on a ground-glass screen.
The optical system of a contour-measuring projector or a comparator requires careful
design and construction. The more important requirements are:
1. The image must be sharp on the screen, and there must be no focusing
errors.
2. There must be no distortion of straight lines into curves.
3. There must be no astigmatism i.e., no unequal definition of the image caused
by horizontal lines being out of focus with vertical lines.
4. There must be no colour fringes.
5. The amount of magnifica-
tion should be capable of
variation.
6. The image on the screen
should not be reversed or
upside down.
The screen is generally placed so
that the image can be read at a dis-
tance of about 300 mm. The screen
must be shielded from sunlight and
other strong lights, and for some work
total darkness is essential. A point to
remember in optical instruments of
this type is that increased magnifica-
tion brings about a reduced filed and
illumination. Thus, too large a mag-
nification should be avoided. Most
industrial instruments have magnifi-
cations ranging from 15x to 50x. The
distortion is less than 0.005 mm per Fig. 5.41 The use of the optical comparator
mm. Even this small distortion does (Jones & Lamson Company)
not affect tests where micrometer
measurements are made on the object itself. The accuracy of the projection system
is important where the shadow contour is checked against a master drawing or tem-
plate. Figure 5.41 shows a typical optical projector.
Figure 5.42 shows the major components of a Jones and Lamson comparator. It will
measure angles in degrees and minutes and is effective in checking lead or spacings,
height, or depth in workpieces and gauges. It is also useful for comparing objects with
a master outline.
The master outline can be drawn to an increased scale on tracing paper and then
placed on the glass screen. Hold-down clips are provided for this purpose. The table
or stage is fitted with micrometers for the measurement of distances traversed.
418 Tool Design

Figure 5.43 shows a line drawing of the optical system of the machine in Fig. 5.42.

Fig. 5.42 Major components of the J & L Comparator (Jones and Lamson Company)

Screen

Screen
Mirror

Condensing
lenses

Lamp Telecentric
Object stop
Projection Basic optical projection system
lens with Telecentric stop

Fig. 5.43 The optical system for Fig. 5.42 (Jones & Lamson Company)

The workpiece or object is positioned so that the magnified contour will fall on the
screen in relation to the correct outline. Figure 5.44 illustrates the optical system of
this machine using reflected light to form the image of the workpiece or object on the
screen.
An example of the use of the optical comparator is shown in Fig. 5.45. The workpiece
is a lever from an adding machine mounted in position for inspection. The shadow
of the lever is projected onto a tolerance chart, which contains two tolerance out-
lines for the part to pass inspection. The advantages of this system of gauging over
Gauges and Gauge Design 419

Reflection attachment

Surface being
projected

Surface illumination

Fig. 5.44 Optical system of machine shown Fig. 5.45 Using a small table-model
in Fig. 5.42 using reflected light optical comparator (Jones &
to form the image of the object Lamson Company)
on the screen (Jones & Lamson
Company)

mechanical gauges are that (i) all elements of the workpiece are checked at a suf-
ficient magnification to show small tolerance errors, and (ii) there is no wear or pres-
sure on the gauge surfaces.

Three-wire system of thread measurement Indicating gauges are often used for
accurate measurements of the pitch diameter of screw threads. The technique used
is referred to as the three-wire system of thread measurement.
The pitch diameter is defined as the diameter of an imaginary cylinder whose sur-
faces pass through the thread at a point where the width of the thread is equal to the
space. This makes an ideal point of measurement because it is least affected by any
variation in the angle of thread. Special wires accurately ground to size to touch the
side of the threads at exactly the right point are known as best-sized wires.
The method of making the measurement is shown in Fig. 5.46, where M is the mea-
surement made over wires, E is the required pitch diameter or effective diameter, q is
one-half the included angle of the thread, G is the diameter of the wires, and p is the
pitch of the thread.
Referring to Fig. 5.47, from geometry,
A = R cosec q
C = R + A = R + R cosec q
= R (1 + cosec q )
p
B = __ cot q
4
The pitch diameter, E can be represented in terms of M, G and p, as
p
E = M – 2C + 2B = M – G (1 + cosec q) + __ cot q ...(5.1)
2
420 Tool Design

Fig. 5.46 Three-wire thread measurement Fig. 5.47 Formula for wire measurement

The equation represented above does not consider the helix angle, and therefore is
a close approximation.
The error due thread angle can be minimised, if the wires make contact at the effec-
tive diameter. In order to reduce the effect of the helix angle, it is important to select a
best wire size for a given combination of thread angle and pitch. The best wire size, Gb
can be derived assuming that the errors in measurements are zero, notwithstanding
the variations of q.
Equation (5.1) may be rewritten as,
p
M – E = G(1 + cosec q ) – __ cot q
2
Differentiating with respect to q, we get
d (M – E)
________ p
= – G cosec q cot q + __ cosec2 q
dq 2
Equating the above equation to zero, the best wire size Gb can be represented as
p
Gb = __ sec q
2
Therefore, form the above equation, it possible to find the best wire size for different
thread forms and for different pitches. Table 5.4 lists the formula for calculating the
best wire size for different thread forms.

Table 5.4 Best wire size for different thread forms


Thread forms Best diameter of measuring wires

Whitworth (55°) 0.5636p


B.A. (47.5°) 0.5463p
Metric (60°) 0.5773p
Unified (60°) 0.5773p

What will be best size of measuring wire diameter for measuring


Example 5.3 a metric thread of 2.5 mm pitch?
Solution: The best size of measuring wire diameter, Gb is related to half the thread
angle, q and pitch, p as:
Gauges and Gauge Design 421

p
Gb = __ sec q
2
For metric thread q = 30°
Given, p = 2.5 mm
Therefore,
2.5
Gb = ___ sec 30° = 1.443 mm
2
What will be the pitch diameter of a metric thread of 2.5 mm
Example 5.4 pitch, if the measurement over the wires (three-wire system) is
12.632 mm?
Solution: Given: p = 3 mm, q = 30° (since thread angle of metric thread is 60°),
M = 12.632
The best size of measuring wire diameter, Gb for metric thread is,
Gb = 0.5773 p = 0.5773 × 2.5 = 1.443 mm
The pitch diameter, E can be represented in terms of M, Gb and p, as
p
E = M – Gb (1 + cosec q) + __ cot q
2
where M is the measurement made over wires, E is the required pitch diameter of
effective diameter, q is one-half the included angle of the thread, G is the diameter of
the wires, and p is the pitch of the thread.
p
E = M – Gb (1 + cosec q) + __ cot q
2
= 12.632 – 1.443 × (1 + cosec 30°) + 1.25 cot 30
= 12.632 – 4.329 + 2.165 = 10.468 mm
Therefore, the pitch diameter of this thread is 10.46 mm.
Figure 5.48 shows the setup of a compara-
tor type gauge to measure the pitch diameter
of the thread plug gauge by the three-wire
method.

Fixturing indicating gauges It is often nec-


essary to design special fixtures on which to
mount various types of indicating heads. The
most common type of fixture is made in the
general shape of a C and is commonly known
as the C-frame fixture. This type is usually
quite satisfactory provided the designer takes
into account several of its weaknesses. The
following paragraphs discuss these weak-
nesses and how they can be prevented.
Fig. 5.48 Visual thread gauge
The general C-fixture consists of heavy base (Sheffield Corp.)
containing provisions for locating the work-
piece to be gauged. An upright member containing a horizontal road for holding the
indicator usually completes the general fixture (see Fig. 5.49a).
422 Tool Design

It is often necessary to construct the fixture with movable joints to facilitate adjust-
ment. Unless it is designed properly, minute movement may occur at these joints.
While this movement may be almost indiscernible at its source, it may be magnified
10 times by the time it is transmitted to the point where the indicator contacts the
workpiece.
Lost motion between components can be greatly reduced if the fixture can be made
in a one-piece unit. However, this generally is not possible; therefore, the joints should
be constructed in a manner that will greatly reduce lost motion. This can be down
by eliminating single locking screws that tend to provide a pivot point around which
microscopic movements take place. Two points should be provided as far apart as
possible (see Fig. 5.49b)
The frame can be tested for lost motion by placing a master or workpiece in position
in the gauge and applying sufficient force by hand to deflect the frame and note the
indicator reading once the force has been removed. Apply force in the opposite direc-
tion and then release it. If there is no significant difference in the two readings, lost
motion is not a factor.

Fig. 5.49 C-frame gauging fixture: (a) General C-frame fixture for mounting indicator heads
(b) Double screws with wide spacing tend to eliminate pivot effect (c) Method of
checking rigidity
Gauges and Gauge Design 423

The C frame should be rigid enough to prevent frame deflection from even the slight-
est pressures. Rigidity of the frame can be increased by locating the extension arm
as low down toward the base on the upright post as possible, with as little overhang
beyond the post as possible. The members themselves should be rigid enough to
withstand normal gauging pressures without deflection.
The rigidity of the frame can be tested by applying a 4.5 N weight on the extension
arm at the centreline of the indicator and noting the amount of deflection on the indica-
tor (see Fig. 5.49c). Then place a precise scale under the indicator and apply pressure
by hand until the indicator again shows the same amount of deflection. Note reading
on the scale. If the ratio of the 4.5 N weight to the scale reading is 100 or more, the
frame is rigid enough for accurate comparison gauging. In other words, the scale
reading should not be more than 0.045 N. This type of test is especially important on
gauges with large frames and excessive overhang.

5.6 AUTOMATIC GAUGES


An automatic gauging machine or fixture is a device that performs one or more of the
following functions: (i) the successive inspection of a number of independent dimen-
sions and the automatic passing of only those workpieces which come within set
tolerances, (ii) the separation of workpieces into a number of classes according to their
actual dimensions, (iii) a combination of the above functions. An automatic gauging
machine may be fed by hand, from a hopper, or by a magazine. The workpieces on
a multistation machine are carried automatically from one gauging station to another.
Workpieces with dimensions within the limits of the tolerances are dropped into a hop-
per at the end of the machine. Solenoid valves can be made to operate chute gates on
the machine to classify workpieces, the actual dimension of the workpiece determin-
ing the chute into which it falls.
A very simple automatic gauge for checking the size of steel balls is shown in
Fig. 5.50. In this, the balls are sent rolling down an inclined raceway fitted with knife-
edges so set that as the correct-sized ball reaches the proper point, it drops into the

A Ball workpieces

Knifeedge
Section A-A

Fig. 5.50 Simple automatic gauge


424 Tool Design

right container. Undersized balls drop through before this point, and oversized balls
roll by and drop through at the end of the raceway.
The tool designer generally is not required to design sophisticated automatic gauging
systems, which is generally done by companies specialising in this type of work with
a background of many years’ experience in the development and manufacture of this
type of equipment. They send skilled applications engineers into the customer’s plant
to survey his products and processes. They then plan, justify, and finally integrate a
system into the customer’s operation.
Sophisticated automatic gauging systems are constructed of components which have
pneumatic, electrical, mechanical or electronic sensing devices and are combined
with various mechanisms for transferring, orienting, sequencing, classifying, accept-
ing, and ejecting workpieces. The resulting systems automatically perform one or
more of the following operations: part detection, part transfer, measurement, classifi-
cation, data storage and recording, segregation, and assembly.
Figure 5.51 and 5.52 show examples of Sheffield automatic gauging machines used
for final inspection. The gauging machine shown in Fig. 5.51 checks eight cylinder

Fig. 5.51 Automatic gauging machine (Sheffield Corp.)

bores simultaneously for size and out-of-round in four planes. Taper (top to bottom) is
computed, and each bore is classified in one of six classes of 0.01 mm. Each bore’s
class is stamped on the block. Machine cycle rate is 2000 per hour.
A unique feature in addition to the general-purpose computer is the control system, or
processor, used in place of conventional electric-relay banks to control sequencing,
Gauges and Gauge Design 425

Fig. 5.52 Automatic gauging machine (Sheffield Corp.)

switching, transfer, etc. Other computer capabilities are electronic mastering without
shutdown and self-diagnosis of malfunction.
Figure 5.52 shows a machine for cylinder-bore gauging and classifying. It has two
pneumatic gauge stations that swing out for greater convenience in setup and ser-
vice. The gauge automatically measures and classifies into nine classes of 0.01 mm
per class and stamps each bore’s class on the block with fast-drying ink. It has a cycle
rate of 150 blocks per hour and is designed to handle two different bores at random.
Out-of-tolerance horse are not stamped, and the machine initiates a signal to eject
the block from the line.

Solved Design Problem

Problem 1 Design a form gauge to check the angles of the workpiece shown in Fig.
5.53 (a) and (c).
Solution: In this design problem, a form gauge is to be designed to check the angles
60° and 135°. A form gauge can be used twice to check each angle.
Figure 5.54 steel, stainless steel or other gauge making material. Both the top and the
bottom susrfaces of the gauge preferably are ground and lapped.
The lower side of the gauge can be used for checking the 60° angle as shown in
Fig. 5.55. For checking tha angle the gauge surface are to be aligned and moved
sidewise.
426 Tool Design

Fig. 5.53 (a) Solid model of the part (b) Dimensioned isometric view
Gauges and Gauge Design 427

Fig. 5.54 Gauge for checking the angles of the workpiece: (a) Solid model (b) Dimensioned
front view of the gauge

Fig. 5.55 The designed gauge for checking the 60° angle on the workpiece
428 Tool Design

Fig. 5.56 The designed gauge for checking the 135° angle on the workpiece

The upper side of the gauge can be used for 135° angle as shown in Fig. 5.56. For
checking the angle the gauge surfaces are to be aligned and moved sidewise.

Summary

Modern manufacturing requires extensive use of gauges for shop work. A gauge is
defined as a device for investigating the dimensional fitness of a part for a specified
function. Gauging controls the dimensions of a product within prescribed limits and
it also helps in segregating the prescribed limits and it also helps in segregating the
products that are not within the limit. Gauges and gauging, therefore is an important
aspect of tool engineering. This chapter delved on the various gauges, gauging pro-
cess and gauge gauging process and gauge design.
• Although standard gauges and gauging methods are available, many a times
specially designed gauges are required for checking dimensions.
• Fixed limit gauges are used to determine wither the product is within the pre-
scribed limit and are intended for use as inspection gauges.
• Ring gauges are used to measure cylindrical surfaces, taper on shafts, and
similar workpieces, as well as external threads.
• Plug gauges are used for checking the uniformity of holes.
• Snap gauges are used for checking diameter and thickness.
• Form gauges are used for checking the contour of a workpiece.
• Length dimensions used in industry are based on the standard unit of dimen-
sional measurement. Gauge blocks have emerged as secondary standard.
• Gauge tolerance is held up 10% of the product total tolerance.
Gauges and Gauge Design 429

• Selection of gauge material is based is based on the tolerance to be checked,


the number of items to be gauged, the composition and hardness of the mate-
rial to be gauged, the cost of the gauge material, and the complexity of the
workpiece to be gauged.

Questions
1. What are the seven basic elements of workpiece geometry that gauges are
designed to check?
2. What is the major difference between fixed-limit gauges and indicating
gauges?
3. What is a fixed master gauge?
4. What are the advantages and disadvantages of fixed-limit gauges compared to
indicating gauges?
5. What is the difference between master gauge blocks, inspection gauge blocks,
and working gauge blocks? How are they used?
6. Why should steel gauge blocks not be wrung together any longer than
necessary?
7. What is monochromatic light?
8. What is meant by the rule of 10 to 1?
9. What is bilateral tolerancing (sometimes called the split-tolerance system)?
10. What is unilateral tolerancing?
11. How are unilateral tolerances applied to (a) plug gauges and (b) ring gauges?
12. What are the important properties of materials from which gauges are
manufactured?
13. How is the dimensional stability of carbon steels and alloy steels used in gauge
manufacture increased?
14. How does deep freezing increase the gauge stability?
15. What are the disadvantages of gauges made from tungsten carbide
materials?
16. Why are indicating gauging methods often referred to as comparative gauging
methods?
17. What are the advantage and disadvantages of the dial indicator as a magnifi-
cation movement?
18. Explain the principle of the reed movement.
19. What are the advantages of the twisted-strip mechanism?
20. Explain the principle of pneumatic gauging.
21. Why does gauging-head war not affect the accuracy of pneumatic gauging?
22. What are the two major types of air-gauging systems?
23. What is the major advantage of electronic gauges?
24. Define the pitch diameter of a thread.
430 Tool Design

25. Why is it necessary to determine the best sized wires when measuring the
pitch diameter of the thread by the three-wire method?
26. How is the frame of a C-frame indicating fixture tested for lost motion?
27. What are the major problems in using the C-frame to position gauge heads?

Problems
1. What is the height for a 125 mm sine bar in order to measure a taper of
27°32¢?
2. A taper is being checked with a 250 mm since bar. The perpendicular is 76 mm.
What is the angle?
3. Using Table 5.2, select the necessary gauge blocks to check dimension of
1.1465.
4. What will be the pitch diameter of a metre thread of 2.5 mm pitch if the
measurement over the wires (three-wire system) is 13.1 mm? Sketch the
thread, wires, and micrometer anvils. Put in all dimensions. Answer must be
accurate to within four decimal places.
5. The basic pitch diameter of a M20 × 2.5 mm thread is 17.412 mm. What will be
the measurement over three wires?

Design Problems
1. Design a plug gauge to check the holes in one of the workpieces shown in
Fig. 5.57a. Use unilateral gauging tolerances.
2. Design a flush-pin gauge to check the depth of the slot in the workpiece shown
in Fig. 5.57b.
3. Design a solid progressive snap gauge to check the diameter of the groove
roots on the workpiece shown in Fig. 5.58. Specify surfaces that are to be
ground and lapped. Use unilateral gauging tolerances.
4. Design an air-gauging application to check the taper of the workpiece shown in
Fig. 5.57c.
5. Design a transparent tolerance chart that can be clipped to the screen of a
5x optical comparator for checking the workpiece in Fig. 5.57d. Show tolerance
outlines on the chart. Assume tolerances for all dimensions on workpiece to
be ± 0.010.
6. Design a template gauge for the workpiece shown in Fig. 5.58. Dimension
fully.
7. Figure 5.59 shows a workpiece with a hole bored at an angle of 10° to the hori-
zontal. Design a fixture to check this angle as suggested in the dotted outline.
8. Design a fixture to check the face angle (61°43¢) and mounting distance (37 mm
of the gear blank shown in Fig. 5.60a. Use a surface plate, a height gauge and
indicator, a sine bar, an arbor, and gauge blocks. Select suitable buttons, and
set the sine bar so that the face F is horizontal.
9. Design a gauge as outlined in Fig. 5.60b to determine the radius of a chord-
shaped workpiece, the radii to range from 660 to 1320 mm. From equations
dealing with the chords of circles, derive a simple formula for obtaining R from
L and M.
Gauges and Gauge Design 431

Fig. 5.57 Workpieces to be checked

Fig. 5.58 Turned workpiece


432 Tool Design

Fig. 5.59 Gauge-design problem: Workpiece


Gauges and Gauge Design 433

Fig. 5.60 Gauge-design problems: (a) Bevel gear blank (b) Radius gauge

References
Books
• ASTME: “Tool Engineers Handbook,” 2nd ed. F W Wilson (ed.), McGraw-Hill, New York, 1959.
• Farago, F T: “Handbook of Dimensional Measurement,” The Industrial Press, New York, 1968.
• IBM: “Precision Measurement in the Metal Working Industry,” Syracuse University Press,
Syracuse, NY, 1952.
• Kennedy, C W, and D E Andrews: “Inspection and Gaging,” 4th ed., The Industrial Press, New
York, 1967.
• Moore, R F: “Foundations of Mechanical Accuracy,” The Moore Special Tool Company,
Bridgeport, Conn., 1970.
• SME: “Functional Inspection Techni1ues,” Dearborn, Mich., 1967.
—: “Functional Gaging of Positionally Toleranced Parts,” Dearborn, Mich., 1964.
—: “Handbook of Industrial Metrology,” Dearborn, Mich., 1967.
Catalogues
• AA Gage Division, US Industries, Detroit.
• John Bath and Company, Worcester, Mass.
• Bausch and Lomb Optical Co., Rochester, NY.
• The Bendix Corporation, Automatic Measurement Division, Dayton, Ohio.
• Brown & Sharpe, North Kingstown, RI.
• The DoAll Company, Des Plaines, Ill.
• Federal Products Corp., Providences, RI.
• Jones & Lamson Company, Springfield, Vt.
• Pratt & Whitney, Cutting Tool and Gage Division, West Hartford, Conn.
• Scherr Tumico, St James, Minn.
• LS Starrett Company, Athol, Mass.
• Taft-Peirce Mfg Co., Woonsocket, RI.
• The Van Keuren Co., Watertown, Mass.
6 LOCATING AND CLAMPING
METHODS

6.1 INTRODUCTION
A large portion of the tool designer’s time is spent in the solution of locating and
clamping problems. The term, ‘locating’ as used in the language of the tool designer,
refers to the dimensional and positional relationship between the workpiece and the
cutting tool used on the machine. A work-holding and positioning device is usually
employed to establish this relationship.
The various forces acting upon the workpiece during a machine operation necessitate
a means of clamping it in position after it has been correctly located. In theory, clamp-
ing and locating are individual and distinct problems, but in practise they cannot be
separated. Many a times, locating and clamping devices are incorporated in the same
mechanism. The method of clamping may depend upon the type of locating device
and vice versa. It should be remembered, however, that the selection of locating and
clamping methods and devices will depend upon the original machining operation and
the configuration of the part.
This chapter will be connected with the basic principles of locating and clamping. Note
that it precedes the chapters on jig design and fixture design. The tool designer must
have a thorough understanding of clamping and locating before attempting the solu-
tion of jig and fixture design problems.

6.2 THE BASIC PRINCIPLES OF LOCATION


The term, ‘location’ refers to the method of establishing correct relative position of
the workpiece with respect to the cutting tool. In order to decide upon the location
method, one has to consider the workpiece shape, surfaces and features that are
likely to obstruct the tool movement or access direction. The correct positioning of the
workpiece essentially requires restricting all the degrees of freedom of the workpiece
essentially requires restricting all the degrees of freedom of the workpiece positively.
This is done with the help of locators, which must be strong enough to resist the cut-
ting forces while maintaining the position of the workpiece.
Apart from correct position of the workpiece, locating devices can be used for fool
proofing. Fool proofing ensures that the component is placed in the jig or fixture in
the correct position and orientation. It does so by restricting the placement of the
workpiece in any other position or orientation. Fool proofing is achieved using fool
proofing pins.
The proper location of a workpiece reqiures use of minimum number of locating
points.
Redundent location is not desired in a jig or a fixture design. For external plain surface
3-2-1 principle of location is generally employed.
Locating and Clamping Methods 435

Configuration is the major factor in determining how a workpiece will be located.


Examination of a typical workpiece from the metals industries shows that the configu-
ration is determined by a combination of flat, circular, and irregular surfaces.
A flat surface is one that lies in one plane, and a circular, and irregular surface is one
that is made from the segment of a circle. The inside surface of the circle segment
may be used to form a concave surface. Irregular surfaces are not flat or circular. They
may or may not be geometrically true.
Each of the three preceding surface types may be rough or finished. A finished sur-
face is one that has been produced by a machining operation. A rough surface is char-
acteristic of materials and workpiece before machining, such as casting, hot-rolled
scale, forgings, weldments, etc.
The tool designer should study the workpiece carefully before deciding which surfaces
are best suited for locating. The following subsections explain the basic principles of
locating from a plane, circular, and irregular surface.

6.2.1 Locating Methods


Locating from plane surfaces The proper location of a workpiece requires use of
minimum number of locating points. Redundant location is not desired in a jig or
a fixture design. For external plain surface, 3-2-1 principle of location is generally
employed. In order to understand 3-2-1 principles of location, let us consider a rect-
angular block with all the planes perpendicular to each other. The rectangular block
has 12 degrees of freedom as shown in Fig. 6.1. There are six rotational and six axial
degrees of freedom. Each degree of freedom is indicating in the figure. In order to
properly locate the block, all the 12 degrees of freedom need to be restricted with
suitable locating points. Location of the block can be done using six locating points;
three locating point in the primary locating surface, two locating points in the second-
ary locating surface and one locating point in the ternary locating surface, as shown
in Fig. 6.2. The rectagular block in Fig. 6.2 is deliberately made transparent so that
all the six locators are visible. Placing the primary locating surface of the block on
the three locating pins restricts five degrees of freedom; one linear movement (along
–Z direction, 6) and four rotational movements (7, 8, 9 and 10). The two locators

Fig. 6.1 Twelve degrees of freedom of Fig. 6.2 Six point location of a rectangular
rectangular block block
436 Tool Design

placed on the secondary locating surface


restrict three more degrees of freedom; one lin-
ear movement (along + X direction, 2) and two
rotational movements (11 and 12). The sixth
locator positioned on the ternary surface of the
block restricts one linear movbement (along +
Y direction, 4). Therefore, six locators restrict
nine degrees of freedom. Remaining three
degrees of freedom is restricted by clamping
the block. The implementation of this principle
is explained in Figs. 6.3–6.7.
The basic reference for locating is a flat plane,
generally a machine table. The machine table
is usually at right angles or parallel with the
machine’s feed movements, as shown in Fig.
Fig. 6.3 Three standard movements
6.3. Most machines have three standard move-
of machine tool table (Cin-
ments which move the workpiece up and down
cinnati Milacron, Inc.)
(vertical), right or left (longitudinal), and in or
out (cross). Bear in mind that the machine being referred to is used in a general
sense, as the feed movements of specific machine tools may vary somewhat.
All locating devices are made with regard to the basic reference plane (machine
table). If a workpiece with a plane surface is placed upon the basic reference plane
(machine table), it will remain in position because of the forces of gravity acting upon
it. However, during a machine operation, the workpiece may move in any direction
except toward the surface of the machine table. The machine table acts as a stop and
becomes a locating surface. It prevents movement of the workpiece when forces are
imposed upon it by the vertical feed movement (see Fig. 6.4).

Cross-feed
Workpiece
movement

Longitudinal feed moveement

Basic reference
plane

Vertical-feed
Machine table
movement

Fig. 6.4 Workpiece may move in any direction except toward the surface of the machine
table
Locating and Clamping Methods 437

The workpiece shown in Fig. 6.4 has a flat side against the machine table. If the work-
piece does not have a flat side to mate with the machine table, as in Fig. 6.5, the flat
plane of the machine table cannot be used as a locating surface. It becomes neces-
sary to reduce the contact area of the locating surface. A series of sharp points would
give the theoretical minimum amount of contact area, but in practise the points must
have enough body to prevent breaking and rapid wear.
A minimum of three points (or locators) must be used to locate the workpiece shown
in Fig. 6.5, although four or more may be used to provide adequate support. It should
be noted that a minimum of three locators will always theoretically establish the same
location of the workpiece. Four or more locators do not establish this theoretical loca-
tion because if the workpiece surface is not true, the position is still determined by
three locators. This may be compared with two tables, one with three legs and one
with four legs. The table with three legs will rest upon all legs on an uneven floor, while
the table with four legs always has one leg that does not contact the floor. The table
is allowed to tilt from one leg to another, and the result is that the table top is in two
different planes.

Workpiece

Basic reference
plane
Locators

Machine table

Fig. 6.5 The use of three locators to establish the workpiece in a plane parallel to the refer-
ence plane

At this point, the base plane is serving as a locator and prevents workpiece movement
when forces are imposed upon it by the vertical feed movement. Forces imposed by
the longitudinal or cross-feed movement will cause the part to shift. Providing stops
on two sides of the part, as shown in Fig. 6.6, will prevent movement of the part. It
should be mentioned, however, that the stops do not locate the part in the same posi-
tion each time the workpiece is place against them. If the workpiece is to be located
in the same position each time, two stops or locators must be provided on one side of
the part, as shown in Fig. 6.7.
Longitudinal-feed
movement

Workpiece Cross-feed
movement

Fig. 6.6 Stops prevent workpiece movement but do not locate the part in the same position
each time it is placed against them
438 Tool Design

Longitudinal-feed
movement

Workpiece Cross-feed
movement

Fig. 6.7 Location by use of three points in the reference plane, two points in a second plane,
and one point in a third plane

The plane-sided workpiece is now completely located although it has not yet been
clamped (clamping will be discussed in Sec. 6.4). Each time it is removed, it can be
replaced in the same position. Complete location has been accomplished by use of
three points in (or parallel with) the reference plane, two points in a second plane, and
one point in a third plane.
The locating points should be placed as far apart as possible to minimise the effect
in inaccuracies in the workpiece and locators, as shown in Fig. 6.8. This separation
also provides greater support because the locators are nearer the outer extremities
of the workpiece.

12.71 9.53

12.71 9.53

Fig. 6.8 Locators should be spaced as far apart as possible to minimise the effect of inac-
curacies (dimensions are greatly exaggerated)

Locating from circular surfaces Here again, the basic reference for location is the
flat plane of the machine-tool table surface. However, instead of locating the flat plane
of the workpiece parallel to the reference plane, it is necessary to locate the axis of
the circular workpiece (see Fig. 6.9).

Fig. 6.9 Circular workpiece must be located with its axis parallel
with the basic reference plane
Locating and Clamping Methods 439

One of the common methods of locating from a circular surface is by using cones, a
method commonly referred to as conical location and usually employed when locating
is done from a hole. The same system may be used when locating on the outside of a
circle, except that the cones are inverted to form cups (see Fig. 6.10).

Workpiece

Machine table Machine table

(a) (b)

Round
workpiece

(c) (d)

Workpiece Clamping force


Locating
cones

(e) Springs

Fig. 6.10 Methods of conical location

Closely related to conical location is the V method, used primarily to locate round
workpieces or workpieces with convex circular surfaces (see Fig. 6.11). It has been
found that the best general V angle is 90°. Smaller included angles hold a round
workpiece more securely but are more susceptible to location errors cause by burrs,
chips, dirt, and workpiece inaccuracies. It should also be noted that any variation in
work size will cause the axis of the round workpiece to shift, as shown in Fig. 6.12.
The V should be directed in such a way that variations in workpiece size will not affect
location on the workpiece.
A method of locating from a circular surface is so common that it is often overlooked
in the use of standard chucks. The method incorporates both locating and clamping
and is often referred to as concentric location because the work is usually positioned
to common centre. Chucks are usually thought of as clamping and locating devices
to be used where both the chuck and workpiece revolve; however, they may be quite
useful for locating and clamping stationary workpieces to machine tables (see Fig.
6.13). The location of a round workpiece is not affected by variations in diameter, and
location and clamping can be accomplished quickly and readily.
440 Tool Design

90°

Workpiece

15°

(a) (b) (c)

Locking
nuts

(d) (e)

Fig. 6.11 Methods of V location

Cutting
Cutting tool
tool

(a) (b)

Fig. 6.12 V should be directed in such a way that variations in


workpiece size will not affect location

Cutting tool

Workpiece

Machine table

Fig. 6.13 Use of a three-jaw universal chuck as a work-locating and clamping device
Locating and Clamping Methods 441

Locating from irregular surfaces Irregular surfaces are neither flat nor circular.
They may or may not be geometrically true. For example, a parabolic or elliptical sur-
face would be considered an irregular surface. A surface of a workpiece that may vary
dimensionally from time to time would also be an irregular surface. An example would
be the raw edge or surface of a casting.
The degree of roughness may determine whether a surface would be considered flat,
circular, or irregular. A rough flat surface may have to be considered as an irregular
surface when determining locating methods, especially when workpiece dimensions
vary from part to part.
Locating methods used for flat and circular surfaces may be used for some irregular
surfaces that are finished and geometrically true. For example, V locating methods
may be used to locate certain parabolic surfaces, and button locators may be used to
locate certain elliptical surfaces.
For the most part, the tool designer will be concerned with location of irregular sur-
faces that vary dimensionally from time to time. Castings, forgings, and weldments are
typical examples. It may be necessary to locate from an irregular surface only during
the first machining operation, as it should produce holes or surfaces that can be used
as reference or locating points for subsequent operations.
The unevenness of the surface of a casting will allow a maximum of three contact
points. More than three points will allow the casting to teeter. Moreover, if more than
three points are used, the workpiece will deform when clamping pressure is applied. It
is therefore necessary to use adjustable rest pins or equaliser to compensate for the
unevenness of the workpiece surface.
One of the simplest types of adjustable rest pins, commonly called a fixture jack, is
shown in Fig. 6.14. In use, the workpiece is positioned on three non-adjustable loca-
tors, and the jack or jacks are adjusted until they touch the workpiece surface. The
contact pressure between the jacks and the workpiece depends upon the judgment of
the operator and is a disadvantage when a number of jacks are used.
The type of jack shown in Fig. 6.14 is suitable only if it can easily be reached by the
operator. Figure 6.15 shows various examples of jacks that can be placed well under
the workpiece.

(a) (b) (c)

Fig. 6.14 Types of adjustable rest pins or fixture jacks

Figures 6.16 and 6.17 show mechanical equalising devices that are sometimes used.
They give nearly equal contact pressure between the two locators and are used to
replace one of three supporting points on a rough workpiece surface. In other words,
442 Tool Design

four contact points may be used, yet the three-contact-point principal is still adhered
to.

Spring

Spring
(a) (b)
Dust cap
Hardened
bushing

(c)

Fig. 6.15 Fixture jacks that can be placed well under the workpiece

Rest pins

Workpiece

Fig. 6.16 Mechanical equalising jacks

Slot for removal


of pin

Rocker
Workpiece

Cone locator pins

Fig. 6.17 Equalising rocker provides an extra point for better workpiece support
Locating and Clamping Methods 443

The preceding discussion of locating to irregular surfaces has been limited to provid-
ing support for the workpiece or locating in one plane. Location in the two remaining
planes is accomplished in a similar manner. The locators are usually adjustable to
allow for dimensional irregularity and may not have locking devices, as shown in Fig.
6.18. The positional location of the part is sometimes determined by sight location,
which is particularly applicable to first-operation location of workpieces with irregular
surfaces. Punch marks, scribed lines, or sight holes are provided in the jig or fixture
and are used to locate the part roughly (see Figs 6.19 and 6.20).

Socket head
set screws
Rough
casting

Brass (b)
insert

(a) (c)

Fig. 6.18 Adjustable locators that can be locked position

Fig. 6.19 Use of sight locators to position rough casting


444 Tool Design

Scribed Rough
lines forging

Adjusting
screws

Sight holes

Fig 6.20 Location of rough forging using scribed lines and sight holes

6.3 LOCATING METHODS AND DEVICES


Section 6.2 was concerned with the basic principles of location. Locating methods dis-
cussed were general, and specific types were not covered. This section is concerned
with specific locating methods and devices.
Through the years, tool designers have developed locating methods and devices that
have proved efficient and simple. The majority of workpieces can be located with
these methods, and the result has been that they have been adopted throughout the
country. The demand for these universally used locators has been great enough to
cause independent tool companies to manufacture them in standard sizes and large
quantities for resale. The cost of the mass produced locator is much cheaper than if
the toolmaker were to make it. For the reason, the tool designer should select locat-
ing methods that require standard, readily purchasable locating devices. Sizes and
dimensions can readily be obtained from tool-component catalogues. Many compa-
nies stock quantities of standard tool components and require that the tool designers
specify only stock items for their designs. Nonstock items are not used except as a
last resort.
The locating devices discussed in the following subsection are for the most part stan-
dard items, which can be found in various tool-component catalogues. Sizes and
dimensions can be determined from these sources.

Pin and button locators One of the common methods of location is to use pins and
buttons. A round pin or button is used to firmly support or hold the workpiece in posi-
tion. The main difference between pins and buttons is in length. Buttons are generally
shorter than pins and are used for vertical location. Pins are usually used for horizon-
tal location. Larger sizes are sometimes referred to as plugs.
Figure 6.21 shows the common methods of utilising pins and buttons as locators. In
Fig. 6.21 a, the button is used for workpiece support (vertical location). In b and c, the
Locating and Clamping Methods 445

button and pin are used to locate the edge of the workpiece (horizontal location). The
pin in d is used to locate from a hole produced by a previous operation.

Workpiece

(a) (b)
Workpiece Workpiece

(c) (d)

Fig. 6.21 Use of pins and buttons as locators

The simplest form of pin locator is a short length of steel drill rod pressed into a hole
in a jig or fixture base. However, since buttons and pins are available commercially as
standard parts, such an arrangement would be used only as a last resort. Standard
pins and buttons are shown in Fig. 6.22. The rest button in a is used when the work-
piece surface is flat, while the spherical-radius button in b is used when the workpiece
surface is rough and irregular. The screw rest button in c is used when a through hole
is not possible and therefore provides a means of easily removing the button. The
pin in e serves as a workpiece support as well as a locating pin, while the button in f
contains a dirt groove when installed in a jig or fixture baseplate.

(a) (b) (c) (d) (e) (f)

Fig. 6.22 Standard buttons and pins


446 Tool Design

Lock-screw pins are sometimes used


where the pin is subjected to excessive
wear and must be frequently replaced Lock screw Locating pin
(see Fig. 6.23). A hardened linear is lock-screw type
pressed into the base of the jig or fix-
ture to receive the lock-screw locating
pin, which in turn is held into the liner
by a lock screw.
Locating pin
Workpieces may use previously located liner
and drilled holes for locating and refer-
ence points. Two pins inserted into the
holes will establish complete horizontal Tapped hole Jig or fixture
base
location, as shown in Fig. 6.24. Any
dimensional variation between centre-
to-centre distance of the holes will
cause the workpiece to bind if round
pins are used. The use of pins of dif-
ferent lengths will overcome binding to
a certain extent, but a more common
method is to use one round pin and one Fig. 6.23 Lock-screw locating pins
diamond pin. Diamond pins are relieved
on two sides to allow for variations in the centre-to-centre distance of the holes. They
locate at right angles to the axis between pin centres, as shown in Fig. 6.24, and
should be placed in the largest hole. Standard diamond pins are shown in Fig. 6.25.

Diamond pin located at


right angles to axis
between pins

Round pin Diamond pin

Fig. 6.24 Use of one round and one diamond locating pin

Large pins used in holes for locating are generally referred to as plugs. As with pins,
close-fitting workpieces have a tendency to stick on plugs. This problem can be elimi-
nated somewhat by using extra clearance between the workpiece and the plugs, but
this practise is limited by workpiece tolerance. Other methods of remedying this situ-
ation are shown in Fig. 6.26.
Locating and Clamping Methods 447

In Fig. 6.26a, a plug of diameter d extending from


a faceplate will not stick in the hole of_____
diameter D
of a workpiece if the plug length l = ÷2WC , where
W is the outside diameter of the workpiece con-
centric with the hole and C = D – d.
In Fig. 6.26b, a projection of diameter d on a
workpiece will not stick in a locating hole of diam-
eter D if the hole is relieved or countersunk so
that the length of engagement____ between the hole
and workpiece is E = l – m £ ÷WC .
At 6.26c, a method of reducing the tendency for
workpieces to stick on a locating plug extending
from a faceplate is shown. The plug is relieved
by cutting away three equal segments at a work-
able angle of 15°, which results in m = 0.35d. A
Fig. 6.25 Diamond locating pins
plug cut in this manner will not stick if its length
is l = 2.4(2a + 0.85d) (D – d). The disadvantage is that a workpiece can be displaced
in three directions on a relieved plug, and the extra error, in inches, of location as
compared with a full round plug is 0.207(D – d), or about 20 per cent less in locating
accuracy.

Faceplate
Workpiece Locator l
Workpiece
a a

D
d DW d W

a
a
m
l Plug
(a) (b)

d
30º m B A
d
V D V D
s
a
l1
l
(c) (d)

Fig. 6.26 Design of nonsticking plugs and locators [From ASTME, “Tool Engineers Hand-
book, “2nd ed., F W Wilson (ed.), McGraw-Hill, New York, 1959, by permission]

An aligning groove may be machined on the end of a plug of diameter d of unlimited


length l to keep it from sticking when inserted in a hole of diameter D (see Fig. 6.26d).
448 Tool Design
_____
Workable dimensions for such a design are l1 = ÷2AC2 , l = md, and B = 0.985d, where
C2, the clearance between the plug pilot and hole equals D – A and m is the coefficient
of friction between plug and hole surfaces, usually about 0.15 to 0.1 for steel. The pilot
diameter may be made as convenient between size d and A = (2d 2/D) – d.*
Spherical locating plugs, as shown in Fig. 6.27, will also help prevent sticking because
of misalignment. This may be explained by the fact that a sphere has only one dimen-
sion and therefore will not bind in a hole no matter which way it may be turned.

Spherical locating
Workpiece plug

Fixture base

Fig. 6.27 Using spherical locating plugs prevents workpiece from sticking

Rest pads and plates Rest pads and plates are used to support and locate work
vertically in a manner similar to rest buttons, but they are used with larger and heavier
workpieces. Rest pads are similar to rest buttons but do not have a shank. They are
held to the jig or fixture baseplate by socket-head cap screws, as shown in Fig. 6.28.

Workpiece Workpiece

Fig. 6.28 Using rest pads and plates for work support and positioning

* These formulas are from ASTME, “Tool Engineers Handbook,: 2d ed., F W Wilson (ed.), McGraw-Hill,
New York, 1959.
Locating and Clamping Methods 449

Rest plates are essentially rest pads with larger bearing surface. They are used to
support larger parts on previously machined surface. Because of the larger bearing
surface, rest plates are sometimes grooved to reduce bearing area and still provide
support rigidity, as shown in Fig. 6.29. This provides a space for small chips that would
affect the accuracy of locating the workpiece. The edges of the groove act as a chip
cleaner to scrape the chips from locating surfaces when the workpiece is slid over
them.

15°

Fig. 6.29 Rest plate containing angular grooves for chip clearance

Nest or cavity location The term, nest


is loosely used by tool designers and tool- Nest or cavity
makers when referring to the area of the
jig or fixture that receives the workpiece. In
the true sense, the nest of a jig or fixture
should be a pocket or cavity into which the
workpiece is placed and located, as shown Workpiece
in Fig. 6.30. Nests may be milled directly
into a baseplate to the same contour as the
workpiece. For blanked parts a nest may be
pierced directly into a piece of sheet stock Area relief Baseplate
and, with a small amount of relieving with a
file, be bolted directly to the baseplate (see Fig. 6.30 Nest or cavity location
Fig. 6.31).
Properly designed nests should locate the workpiece without supplementary locating
devices, although locating pins are occasionally used. The workpiece should protrude
above the nest far enough to allow the operator to remove it easily. For a workpiece
like that in Fig. 6.31 finger holes or clearance must be provided. It may be necessary
to provide relief under the part to allow for chips that would prevent the workpiece
from seating properly (see Fig. 6.30). Nests may or may not require clamping devices,
depending upon the direction of cutting forces.
450 Tool Design

Finger relief

Workpiece

Sheet stock

Baseplate

Fig. 6.31 Nest formed by piercing the workpiece shape into sheet stock

One disadvantage of nest location is that the workpiece is often difficult to lift out of
the cavity because it is completely surrounded. This can be overcome by providing
an ejecting device, but this adds additional expenses. The use of a partial nest, as
shown in Fig. 6.32, also permits easier removal of the part. Partial nests are often
easier to construct because the entire contour of the workpiece does not have to be
machined.

Workpiece

Chip relief

Fig. 6.32 Use of partial/nest

Locating methods and chips control Machining metal involves the removal of waste
material in the form of chips. Large chips can be removed quickly from the work loca-
tion area by various means, but small chips tend to collect in corners and other inac-
cessible areas and cause error in location. The tool designer should keep this problem
Locating and Clamping Methods 451

in mind when selecting locating devices and either use locating methods that are not
susceptible to error caused by chips or design chip-control features into them.
Chip control at locating points in general consists of providing relief or opening at
areas where chips would tend to accumulate and prevent the workpiece from properly
contacting the locator. Figure 6.33 shows various methods of chip control of location
points. At a, excess material is removed to reduce the contact area between the sup-
porting surface and the workpiece and is often referred to as area relief. This provides
less contact area and an opening where small chips may accumulate without interfer-
ing with the location of the workpiece.
Figure 6.33b shows a chip or burr groove around a locating pin. A similar arrangement
is shown at c, where the groove is provided by the construction of the locating pin.
In Fig. 6.33d, e, and, various methods of providing relief in corners are shown. Note
the open construction of the locator at e to allow chips to pass under the locator with-
out becoming packed in corners.

Workpiece

Workpiece

Area relief Chip and burr


groove
(a) (b)

Chip groove

(c) (d)

K K

K-K

J J

J-J

(e) (f)

Fig. 6.33 Methods for controlling chips and burrs


452 Tool Design

6.4 THE BASIC PRINCIPLES OF CLAMPING


Once a workpiece is located, it is necessary to press it against the locating surface
and hold it there against the forces acting upon it. The tool designer refers to this
action as clamping and the mechanisms used for this action are known as clamps.
Numerous types of clamps have been developed, and it is sometimes difficult to
choose between two types when designing a specific tool. In general, the choice of
clamp is largely determined by the workpiece and the kind of operation to which it is
applied. Simple workpieces of uniform contour need only simple clamps, but casting,
forgings, and other workpiece with irregular or variable contour require very careful
design of the clamping device.
However, simple or complex, all champs must fulfill four essential requirements:
(i) The workpiece must be held rigidly while the cutting tools are in operation; (ii) The
time required for loading or unloading the tool must be as short as possible, which
means the clamping device must be quick-acting; (iii) When subjected to vibration,
chatter, or heavy pressure, the clamping must be positive; and (iv) The clamp must
not damage the workpiece.
The frequency of set-ups will influence the method of clamping. Where large numbers
of workpiece are concerned, quick-acting hydraulic or pneumatic claming devices
may be justified. Clamps that move during tightening should be avoided; however,
certain clamp design that tend to move the workpiece against the locator or into the
nest may be advantageous.
In general, simple clamping mechanisms are preferred to complicated ones.
Complicated clamps with several moving parts are difficult to maintain and tend to
lose their effectiveness as they wear. Clamp components that are subject to wear
should be designed in such a way that they are easily replaced. Such parts should
also be heat-treated to give them maximum wear resistance.
Clamping forces should be directed toward the nest or locating surfaces. They should
be arranged in such a manner that the thrust of the cutting tool is away from the clamp.
Clamping forces should be applied over a heavy part of the workpiece whenever pos-
sible, preferably directly over a rest pad or jack in order to prevent distortion of the
workpiece. Clamping mechanisms should be designed in such a manner that they
cannot be applied in any way except the correct one.
The type of clamp will depend upon the kind of operation to which it is applied. For
example, a milling operation will require a strong clamp strategically located because
cutting forces may change directions as the cutter enters and leaves the workpiece.
Also, a milling operation induces vibrations in the workpiece, and the clamps must be
of a construction that will not loosen under vibration.

Types of clamps While the types of clamps are numerous, they can be classified in
seven basic group: strap, cam, screw, latch, wedge or key, toggle, and rack and pinion.
Most clamping devices contain one or more of these elements. In a combination, the
name of the most prominent element is given to the complete clamping device.
Commonly used clamps are readily available from tool specialty companies. They are
often referred to as standard clamps and should be used whenever possible.
Locating and Clamping Methods 453

Strap clamps The simplest and probably most commonly used clamp is the strap
clamp (Fig. 6.34). All strap clamps employ the principles of levers. Numerous styles
and shapes are available as standard parts to suit different applications.
The essential parts of a strap clamp are shown in Fig. 6.34. Note that the strap (clam-
ing bar) can be slid onto and off of the workpiece for easy workpiece removal. A com-
pression spring between the base and clamping bar holds the bar in position when the
clamping knob is loosened. The pillar, or heel pin, rests in a groove on the underside
of the bar to prevent it from turning when the clamping knob is loosened. A hardened
washer is used next to the clamping knob to prevent it from rubbing the bar.

Slot

Clamping knob
Spherical
washer
Clamping bar

Groove
Workpiece
Pillar pin

Spring

Fig. 6.34 Essential parts of a standard strap clamp

Standard
Angle of maximum spherical
declination washers

Fig. 6.35 Use of spherical washers


454 Tool Design

A two-piece spherical washer is sometimes used to provide proper seating of the nut,
washer, and bar when the workpiece height varies. The upper part of the spherical
washer fits properly under the nut, and the lower part fits on the bar, even though the
stud and bar are not perpendicular to each other (see Fig. 6.35). Spherical washers
are standard parts obtainable from any tooling specialty company.
Strap clamps may be actuated by hand knobs, as shown in Fig. 6.34, or by hex nuts,
cams, hydraulic or pneumatic cylinders, or other devices. The fulcrum of the strap
calm should be located as close as possible to the workpiece in order to obtain maxi-
mum mechanical advantage.
The tool designer is not often required to design and dimension strap clamps because
they are readily available as standard parts, but he may occasionally be required to do
so. The experienced tool designer may rely on his past experience when dimensioning
a simple strap clamp, but as the beginner
lacks experience, he may find the following
information helpful. w

The ASTME “Tool Engineers handbook” d


recommends that the central slot in a typical
strap clamp be 1.6 mm wider than the bolt
or stud passing through it. Common practise h
is to make the width of the strap about the
same as the diameter of the washer under
the clamping nut (see Fig. 6.36). For a stud a b
of nominal diameter d mm, the width of the l
strap may be W = 1.16 + 2.3d. An adequate
____
thickness for the strap is h = 0.92 ÷abd /l. Fig. 6.36 Typical strap clamp [From
If the stud__is in the middle of the span, ASTME, “Tool Engineers
h = 0.46 ÷dl . The proper stud size for the Hand-book,” 2nd ed., F W Wil-
required clamping pressure can be deter- son (ed.), McGraw-Hill, New
mined from Table 6.1. York, 1959, by permission]

Table 6.1 Stud size selection


Clamping Force Required (N) Stud size

2224 M6
4003 M8
6672 M10
9786 M12
17793 M16
28024 M20
40034 M24

Other types of clamps based on the lever principle are shown in Figs. 6.37 to 6.39.
Figure 6.37 shows two types of dog-point clamps, sometimes referred to as toe
clamps. Figure 6.38 shows a pinch clamp. Clamps of this type are used when it is
necessary to clamp against the edge of the workpiece.
Locating and Clamping Methods 455

Workpiece

Workpiece

Fig. 6.37 Two forms of dog-point clamps

Socket-head
cap screw

Workpiece

Fig. 6.38 Pinch clamp

Figure 6.39 shows a quick-acting swinging type of strap clamp. A design of this type is
suitable only for applications where the forces acting against the workpiece are light
and heavy clamping pressure is not required.
456 Tool Design

Workpiece

Section A-A

Fig. 6.39 Quick-acting, swinging-type strap clamp

Cam clamps Properly designed cam clamps provide an effective and rapid means
of clamping but they are not a positive clamping method and may loosen under vibrat-
ing forces. They should not be used where heavy clamping forces are required. Since
they also have a tendency to shift the workpiece or other integral parts of the clamp-
ing device, they can be used to advantage when it is desirable to push the workpiece
against stops or into a nest.
The two main types of cam clamps are spiral and eccentric. Figure 6.40 and 6.41
show the cam contours of each type. An eccentric cam is easier to make but does not
lock as well as a spiral cam.
Generally speaking, the tool designer is not
required to specify and lay out cam contours
P
because cam clamps are readily available E
B
as standard parts from various tool specialty
companies. All that is required is to select the L
F
proper cam clamp from the many listings in tool C A
catalogues and specify the clamp by manufac- r
turer and catalogue number. However, the tool
designer should be able to lay out a cam con-
tour, and the following procedures should help Fig. 6.40 Eccentric-cam contour
the beginner.
The eccentric cam shown in Fig. 6.40 is a direct acting cam with an outline in the
form of a circle of radius r with centre at c. The lever is pivoted by the pin p, and the
eccentricity is E.
Locating and Clamping Methods 457

Suppose the handle is rotated in a clockwise direction. Centre C will travel in a circle
A until the circumference contacts the workpiece at B. Any force applied on the handle
will be exerted at B with an increased magnitude depending upon the leverage L. But
there will be a counteracting force resulting from the force F acting with the lever arm.
If there is sufficient friction and the force is not too large, the clamp may hold, but the
action is not certain.
The spiral cam in Fig. 6.41 is a better outline. The cam curve is a spiral generated as
follows:
1. Determine the rise, which is the distance required for the locking action. The
throw is usually from 60 to 90o.
2. Suppose the rise is the distance ab and that the approximate size of the cam
is determined. Draw the concentric circles A and B, as shown. The throw of the
handle should be about 90o from the completely locked to completely unlocked
position; 10o should be added to each end as a precaution against inaccurate
assembly. This totals 110o. Half this amount, 55o, should be laid off from the
centreline.
3. Draw radial lines x and y.
4. Divide rise ab into the same number of parts.
5. Number each part as shown.
6. Draw an arc from each numbered part on ab to meet its corresponding number
on a radial line, thus establishing points through which a smooth curve can be
drawn to form a true spiral for the working face of the cam.
A study of Fig. 6.41 shows that no matter where the contour of this cam contacts the
workpiece, the surface of the workpiece is tangent to the cam at the centreline and
there can be no lever arm. Consequently, there can be no torque tending to unlock
the clamp.

10°
Workpiece c
5 Handle
y
4

3
Throw CL
2
x
1
B
32 b
10
a
54

Relief
Locating pins 10°
A

Fig. 6.41 Spirial-cam contour


458 Tool Design

Figures 6.42 to 6.46 show various types of cam clamps. Figure 6.42 shows a cam-
clamps assembly in which the cam handle is at the centre of the strap. When this is
not convenient, the assembly of a similar construction (Fig. 6.43) may be used. The
beginning tool designer should find these drawing of value in designing practical jigs
or fixtures.

Fig. 6.42 Cam-clamp assembly

Fig. 6.43 Cam-clamp assembly

Figure 6.44 shows a modified cam and screw clamp for use where space is limited, as
under an overhanging part or workpiece. Figure 6.45 shows a spiral-cam edge clamp
that has the advantage of compactness.
The shaft-eccentric cam clamp is sometimes used for universal jig construction; it
allows high pressure to be exerted and also affords an adequate clamping range or
linear movement. Figure 6.46 shows the outline of a jig with a shaft-eccentric clamp.
Locating and Clamping Methods 459

Fig. 6.44 Cam and screw clamp assembly

Screw clamps Screw clamps incorporate a


screw thread to clamp a workpiece. They have
the advantage of exerting adequate force besides
resisting loosening tendencies set up by vibra-
tion. They are, however, relatively slow and may
not be suitable for high production.
Screw clamps vary from a simple setscrew to
complicated clamping assembly actuated by a
screw thread. Many standard items are available
for the construction of screw clamps, as shown Fig. 6.45 Spiral-cam edge clamp
in Fig. 6.47.

Bolt
Setscrews Handle
Eccentricity

Frame Bushing Bushing

Fig. 6.46 Shaft eccentric clamp

Numerous devices have been developed to speed up the clamping action of a


screw.
460 Tool Design

Fig. 6.47 Standard items used in the construction of screw clamps

Figure 6.48 shows a quick-acting screw device in which a single thread with one turn
is cut in the plunger together with a slot parallel with the centreline. A pin engages in
the slot so as to allow the plunger to slide freely with a rapid movement when the pin
is the groove and to give a powerful thrust to the plunger when it is twisted with the pin
in the thread. Thus, the workpiece can be released with a partial turn of the knob and
the plunger quickly pulled clear for reloading. This device, also called a cum-lock long-
travel clamp, may be classified as having a cam action rather than a screw action.

Fig. 6.48 Quick-acting screw clamp

A quick-acting knob, like that in Fig. 6.49, can be pur- A


chased as a standard hand knob. The conversion is
accomplished by drilling a hole 0.8 mm larger than the
bolt size at an angle of 100 with the axis. Figure 6.49
shows the arrangement of the two intersecting holes,
and it will be observed that while the knob is weak-
ened, the remaining threads have sufficient strength
to withstand moderate loads.
In action, the knob is backed a few turns to release A
the pressure on the clamping bar and then tipped Fig. 6.49 Quick-acting knob
10°, aligning axis AA with the bolt. This disengages
Locating and Clamping Methods 461

the threads, allowing the knob to slip back without being turned. After reloading, the
knob is quickly pushed back along the bold until it is close to the clamp and then
straightened out to the threading position, where a turn or so locks the workpiece into
position.
Figures 6.50 and 6.51 show a screw-clamp assembly often referred to as a swing
clamp. In action, the locking screw is loosened and the clamp assembly is revolved
90° to allow workpiece removal. The assembly in Fig. 6.51 must be inserted well into
the parent fixture in order to provide adequate backing for proper support.

Workpiece

Fig. 6.50 Swing-clamp assembly

Workpiece

Workpiece

Fig. 6.51 Swing-clamp assembly


462 Tool Design

Latch clamps This type of construction is limited to relatively light work, as it is dif-
ficult to secure rigidly. The main advantage is in the ease and speed of manipulation.
Figure 6.52 illustrates a simple form of latch. The leaf is a pivoted member of a jig that
may or may not contain bushings. In operation, the jig is loaded and unloaded upside
down. The leaf is unlatched and then swung open, and the workpiece is replaced by
a new one. Locating pins ensure that the workpiece is placed right. After loading, the
leaf is swung to a closed position, and as surface S strikes surface L, the latch swings
back to allow the leaf to pass. Once the leaf is in a horizontal position, the latch of
forced back by the spring, where it catches and locks the leaf.

Spring Drill bushing

Pressure spring
Pivot

To unlatch pivoted leaf


Jig foot
L S

Fig. 6.52 Simple latch clamp

Pivot Leaf
Figure 6.53 shows a cam-type latch. The working face
should conform to the contour of a spiral cam and should
be generated by one of the methods outlined under cam
clamps. Pin

Figure 6.54 shows a self-locking latch, which is very rapid Cam face
in action. The operator can open the jig with one hand by Jig frame
pushing down on the pin as he raises the leaf.
The common thumbscrew latch is perhaps a little stron- Fig. 6.53 Cam-type latch
ger than other latches. Care must be taken when design-
ing this type to ensure that the operator can turn
the screw easily, even when the threads are Unlock
covered with a certain amount of dried oil, dust, Pin
and chips. The designer must also satisfy him- Jig leaf
self that the toolmaker will be sure to tighten the
screw in the position shown, i.e., perpendicular
to the slot (see Fig. 6.55). This is also known as
a quarter-turn screw.
Pawl
Wedge clamp A plain wedge clamp consists of
a movable inclined plane which forces the work- Spring
piece against a fixed stop (see Fig. 6.56). Plain
Spring
wedge clamps are crude and rarely satisfactory
since they have a tendency to loosen under Fig. 6.54 Self-locking latch
Locating and Clamping Methods 463

vibration. A hammer of some sort is required to drive the wedge in and out, and loose
tools can get lost or mislaid.

Fig. 6.55 Thumbscrew latch (quarter-turn screw) Fig. 6.56 Plain-wedge clamp

A plain wedge clamp can be improved


by adding levers and links (Fig. 6.57), an
arrangement that is quick-acting and pro-
vides a partial lock when the centre pivot
goes past the centreline and the handle hits
the stop pin.
The taper angle of a plain wedge may range
from 6 to 18°, depending upon the coefficient
of friction of the metal. The smaller angle
gives the wedge more holding power but
increases both the distance required to lock
and the forces required to retract the wedge.
The presence of oil and coolants require that
Fig. 6.57 Improved plain-wedge clamp
a smaller angle be used unless an additional
holding device is added.
Combinations of the wedge with other elements reduce some of the objections:
Fig. 6.58 shows a pinch clamp employing the wedge principle in conjunction with a
strap clamp, and Fig. 6.59 shows the principle applied to a pair of clamps to ensure
balanced pressure.

Fig. 6.58 Pinch clamp with wedge Fig. 6.59 Strap clamps employing the
wedge principle
464 Tool Design

Toggle clamps Toggle clamps depend upon the movement of rigid links for their
movement. Figure 6.60 shows a simple toggle arrangement.

Fig. 6.60 Simple toggle arrangement

Several companies have developed standard toggle clamps, and specifications are
readily available from catalogues (see Figs. 6.61 and 6.62). These clamps have been
used entensively to hold sheet-metal parts in position while they are being welded or
otherwise fastened, and they provide relatively heavy clamping pressure, are quick-
acting, and give ample clearance for loading and unloading. An adjustment is usually
provided for variation in stock thickness.

Fig. 6.61 Vertical toggle clamp (Carr Lane Mfg Co.)

Fig. 6.62 Push toggle clamp (Carr Lane Mfg Co.)


Locating and Clamping Methods 465

Rack-and-pinion clamps Rack-and-pinion clamps are used extensively on univer-


sal jigs (see Chap. 7). Since a rack and pinion is not irreversible, (i.e., pressure on the
rack will rotate the pinion), a lock must be incorporated to lock or hold the shaft while
it is being clamped. Universal jigs and rack-and-pinion fixture units produced commer-
cially as standard components contain patented locking machines, usually in the form
of a wedge or can that is actuated by the rack and is nonreleasing.

Hydraulic and pneumatic clamping Many of the previously mentioned clamps may
be actuated by hydraulic or pneumatic methods when large production quantities jus-
tify them. The advantages are faster clamping, uniform and equalised clamping pres-
sure, and less operator fatigue.
Pneumatically operated clamps differ from hydraulically operated ones in the size
of the cylinder, which is smaller with hydraulics because of the higher pressures.
Generally speaking, pneumatically operated clamps are not used where heavy clamp-
ing pressures are required because extremely large cylinders would be required.
Pneumatically operated clamping devices are convenient because most manufactur-
ing firms have a supply of compressed air available. However, the line pressure may
be too low to provide adequate pressure to cylinders for clamping purposes. This con-
dition can be remedied by the installation of an air-to-hydraulic booster, which is used
to convert low pressure air into higher hydraulic pressure for operating a hydraulic
work cylinder. The booster operates from regular shop-line pressure without pumps
or high pressure valving.
The principle of the air-to-hydraulic booster is as follows (Fig. 6.63). If the air-cylinder
piston is subjected to 1 MPa air pressure and the area of the piston is 0.001 m2 a force

Fig. 6.63 Air-to-hydraulic booster circuit


466 Tool Design

of 1 KN is placed upon the ram. If the ram


diameter cross sectional area is 0.0001
m2, the pressure upon the hydraulic oil
must be 10 MPa a hydraulic pressure of
10 MPa has been produced from a 1 MPa
line air pressure, i.e., has increased the
pressure by a ratio of 10:1.
The ratio of an air-to-hydraulic booster can
be determined by dividing the driving air
piston by the area of the driven ram. The
high pressure output is calculated by mul-
tiplying the input-line air pressure by the
ratio between the piston and ram areas.
Air-to-hydraulic boosters can be bought
in a wide range of ratios. Ratio selec-
tion depends upon available air pressure
and clamping-pressure requirements. Fig. 6.64 Typical air-to-hydraulic power
Engineering data are available from pro- booster (De-Sta-Co Division,
ducers of pneumatic products. Figure 6.64 Dover Corporation)
shows a typical air-to-hydraulic power
booster.
Figure 6.65 shows a typical arrangement of hydraulic power clamping from standard
shop-line pressure, using standard components. The retracting clamp, shown at the
left, lifts and retracts to clear the workpiece for easy unloading. At the top, a pressures
point is shown mounted in a straight or 180° bracket. The cast-iron body ram at the
right is being used to actuate a simple strap clamp.

Fig. 6.65 Hydraulic power clamping from standard shop-line pressure (De-Sta-Co Division,
Dover Corporation)
Locating and Clamping Methods 467

Hydraulically operated clamping devices were once limited to use on hydraulic


machine tools. Now, however, self-contained hydraulic clamping systems are avail-
able at a reasonable cost from tooling specialty companies. Such systems of a manu-
ally operated pump, small compact cylinders, standard mounting brakes, a pressure
gauge, and necessary tubing. They are entirely self-contained, and no external power
connections are required; this allows the entire unit to be rotated with the fixture or
moved from one machine tool to another.
The system is fully equalising, as pressure is applied to each clamp evenly. Clamping
pressures can be repeated by observing the pressure gauge during clamping. Rough
workpieces whose dimensions vary from part to part can be clamped because size
variations are automatically compensated for by the length of the clamping cylinder
stroke.
Figures 6.66 and 6.67 show a typical installation of a self-contained hydraulic clamp-
ing system. Figure 6.66 shows the use of a manually operated lever pump designed
to be set up and operated within 30° of the vertical position, and Fig. 6.67 shows a
screw pump that can be used in any position; it may be screwed directly to the fixture
or located remotely.

Fig. 6.66 Self-contained hydraulic clamping system using lever pump


(Vlier Engineering Corporation)

Fig. 6.67 Self-contained hydraulic clamping system using screw pump


(Vlier Engineering Corporation)
468 Tool Design

Solved Design Problem

The design problems discussed in this chapter and subsequent chapters use com-
mercial solid modelling program. The design schemes are presented using solid mod-
els of individual parts and the assembly of the parts. Readers may also use manual or
computer-aided drafting techniques for solving these problems.
Tolerances are not incorporated in the design. Manufacturing of the parts would
require incorporation of tolerances depending upon the available manufacturing facili-
ties. Standard fasteners and parts are used in the design wherever applicable.

Problem 1 Design a strap clamp for holding the workpiece shown in Fig. 6.68
(a) and (b).

Solution: The given problem requires designing a strap clamp to hold a particular
workpiece. Strap clamp can be of various types. A standard strap clamp is designed
for holding this particular workpiece. The designed parts of the strap clamp are pre-
sented in the Table 6.2 with reference to the solid model and detailed drawing of
individual parts.

Fig. 6.68 The finished workpiece (a) Solid model (b) Detailed drawing
Locating and Clamping Methods 469

Table 6.2 Strap clamp parts


Sl. no. Part name/ Part description Quantity Figure Reference

1. Base Plate 1 6.69 (a) and (b)

2. Stud 1 6.70 (a) and (b)

3. Pillar pin 1 6.71 (a) and (b)

4. Compression spring 1 6.72 (a) and (b)

5. Strap 1 6.73 (a) and (b)

6. Spherical washer 1 6.74 (a) and (b)

7. Lever nut 1 6.75 (a) and (b)

Fig. 6.69 Base plate (a) Solid model (b) Detailed drawing

The complete assembly of all the parts is shown in Fig. 6.76. The exploded view of
the assembly is shown in Fig. 6.77. The base plate should be heavy and preferably
made of cast iron. The clamping stud is screwed on to the base plate and the pillar pin
may be press fit on the base plate. The strap can be slid onto and off the workpiece
for easy removal. When the lever nut is loosened, a compression spring between
the base and strap holds the strap in position. The pillar pin rests in a groove on the
underside of the bar to prevent it from turning when clamping lever nut is loosened. A
two-piece spherical washer is used next to the clamping nut. This washer is used for
accommodating the variation of the workpiece height. In order to clamp the strap on
to the workpiece, a knob nut can also be used. However, in this design, a lever nut is
used for faster clamping.
The designed clamp can also be used for workpiece of different sizes (limited to stud
size and spring free and compressed length). However, the pillar pin needs to be
changed according to height of the workpiece to be clamped.
470 Tool Design

Fig. 6.70 Stud (a) Solid model (b) Detailed drawing

Fig. 6.71 Pillar pin (a) Solid model (b) Detailed drawing
Locating and Clamping Methods 471

Fig. 6.72 Compression spring (a) Solid model (b) Detailed drawing

Fig. 6.73 Strap (a) Solid model (b) Detailed drawing

Fig. 6.74 Spherical washer and conical seat (a) Solid model (b) Detailed drawing
472 Tool Design

Fig. 6.75 Lever nut (a) Solid model (b) Detailed drawing

Fig. 6.76 Solid model of the designed strap clamp


Locating and Clamping Methods 473

Fig. 6.77 Exploded view of the designed strap clamp

Problem 2 Design a cam operated strap clamp for holding the same workpiece as
shown in Fig. 6.68 (a) and (b).

Solution: Cam clamps are not intended for heavy clamping or when high vibrations
are expected during processing of the workpiece clamped. Cam clamp in mainly used
for rapid clamping of light workpieces.
The stud used in strap clamp is replaced by an eye bolt in cam clamp; the lever nut
in strap clamp is replaced by a cam lever in cam clamp. Other essential parts of the
designed cam clamp remain same as that of the strap clamp.
The designed parts of the cam clamp are presented in the Table 6.3 with reference to
the solid model and detailed drawing of individual parts.

Table 6.3 Cam clamp parts


Sl. no. Part name/ Part description Quantity Figure Reference

1. Base Plate 1 6.69 (a) and (b)


2. Eye Bolt 1 6.78 (a) and (b)
3. Pillar pin 1 6.71 (a) and (b)
4. Compression spring 1 6.72 (a) and (b)
5. Strap 1 6.73 (a) and (b)
6. Cam Lever 1 6.79 (a) and (b)
7. Screw 1 6.80 (a) and (b)
474 Tool Design

Fig. 6.78 Eye bolt (a) Solid model (b) Detailed drawing

Fig. 6.79 Cam lever (a) Solid model (b) Detailed drawing
Locating and Clamping Methods 475

Fig. 6.80 Screw (a) Solid model (b) Detailed drawing

The complete assembly of all the parts is


shown in Fig. 6.81. The exploded view of
the assembly is shown in Fig. 6.82. The
eyebolt is screwed on to the base plate
and the pillar pin may be press fit on the
base plate. The strap can be slid onto
and off the workpiece for easy removal.
When the cam lever is loosened, a com-
pression spring between the base and
strap holds the strap in position. The pil-
lar pin rests in a groove on the underside
Fig. 6.81 Solid model of the designed cam
of the bar to prevent it from turning. The
clamp
working of a cam clamp is similar to that
of the standard strap clamp, but the clamping and unclamping is much quicker. In
order to unclamp the workpiece, the cam is rotated upwards on the strap, and then
the cam and the eye bolt together is rotated about the vertical axis. This unscrews the
eye bolt upward, releasing the strap from the workpiece.

Fig. 6.82 Exploded view of the designed cam clamp


476 Tool Design

Problem 3 Design a hinged/pivot clamp for holding the workpiece shown in Fig.
6.83 (a) and (b).

Fig. 6.83 Workpiece (a) Solid model (b) Detailed drawing

Solution: The given prob-


lem requires clamping
of the workpiece using
a hinged clamp. Hinged
clamp is also called pivot
clamp. Hinged clamp pro-
vides rapid clearance for
loading and unloading of
the workpiece. The work-
piece is clamped using a
hinged eye bolt, a hinged
strap plate and a floating
pad. The clamp is pro-
vided with open slot on the
hinged strap plate through
which the eye bolt can
be swung in position and
locked in position using a
lever nut. The hinged plate
can be swung aside during
loading and unloading.
The designed parts of
the hinged clamp are
tabulated in the Table 6.4
with reference to the solid
model and detailed draw-
ing of individual parts. Angle base (a) Solid model (b) Detailed drawing
Locating and Clamping Methods 477

Table 6.4 Hinged clamp parts


Sl. no. Part name/ Part description Quantity Figure Reference

1. Angle base 1 6.84 (a) and (b)


2. Eye bolt 1 6.85 (a) and (b)
3. Hinge pin 2 6.86 (a) and (b)
4. Locking pin 3 6.87 (a) and (b)
5. Hinged strap plate 1 6.88 (a) and (b)
6. V shaped floating pad 1 6.89 (a) and (b)
7. V locator pivot pin 1 6.90 (a) and (b)
8. Lever nut 1 6.75 (a) and (b)

Fig. 6.85 Eye bolt (a) Solid model (b) Detailed drawing

The complete assembly of all the parts of the hinged clamp is shown in Fig. 6.91. The
left side and front view of the assembly is shown in Fig 6.92 (a) and (b). The exploded
view of the assembly is shown in Fig. 6.93.
Method of clamping and unclamping In order to clamp the workpiece, it is first
placed on the base. The hinged strap plate is then swung in position over the work-
piece. The pivoted floating pad on the hinged strap plate presses the workpiece in
478 Tool Design

Fig. 6.86 Hinge pin (a) Solid model (b) Detailed drawing

Fig. 6.87 Locking pin (a) Solid model (b) Detailed drawing

required position. The work is pushed against the base which also houses a swinging
eye-bolt. The eye bolt is swung in position in the open slot on the hinged strap plate
and locked in position using a lever nut. In order to unclamp the workpiece, the lever
nut is loosened only by half a turn before the eye bolt and hinged strap plate moved
aside for unloading the workpiece.
Locating and Clamping Methods 479

Fig. 6.88 Hinged strap plate (a) Solid model (b) Detailed drawing
480 Tool Design

Fig. 6.89 V shaped floating pad (a) Solid model (b) Detailed drawing

Fig. 6.90 Floating pad pivot pin (a) Solid model (b) Detailed drawing

Fig. 6.91 Solid model of the designed hinged clamp


Locating and Clamping Methods 481

Fig. 6.92 The designed hinged clamp: (a) Left-side view (b) Front view

Fig. 6.93 Exploded view of the designed hinged clamp

Problem 4 Design a locating and clamping method to drill all the holes in the work-
piece shown in Fig. 6.94 (a) and (b).
482 Tool Design

Fig. 6.94 The finished workpiece (a) Solid model (b) Detailed drawing

Solution: The given problem requires locating and clamping of the workpiece so
that all the five holes can be drilled. This essentially requires designing a drill jig. The
workpiece can be located using a nest or cavity locator and can be clamped by screw
clamping.
The designed parts of the drill jig are presented in the Table 6.5 with reference to the
solid model and detailed drawing of individual parts.

Table 6.5 Designed drill jig parts


Sl. no. Part name/ Part description Quantity Figure reference

1. Base Plate/ Bush plate 1 6.95 (a) and (b)


2. Drill Bush for φ 20 mm Hole 1 6.96 (a) and (b)
3. Drill Bush for φ 14 mm Hole 2 6.97 (a) and (b)
4. Drill Bush for φ 6 mm Hole 2 6.98 (a) and (b)
5. Screw Clamp 4 6.99 (a) and (b)
6. Screw Clamp Bush 4 6.100 (a) and (b)
7. Screw Clamp Liner 4 6.101 (a) and (b)
8. Hex Nuts 4 6.102 (a) and (b)
Locating and Clamping Methods 483

Fig. 6.95 Bush plate (a) Solid model (b) Detailed drawing

Fig. 6.96 Bush 20 mm (a) Solid model (b) Detailed drawing


484 Tool Design

Fig. 6.97 Bush 14 mm (a) Solid model (b) Detailed drawing

Fig. 6.98 Bush 19 mm (a) Solid model (b) Detailed drawing

The complete assembly of all the parts is shown in Fig. 6.103. The exploded view of
the assembly is shown in Fig. 6.104. The bush plate is first tuned over and workpiece
is placed in the cavity and clamped in position. The drill jig is then rotated and placed
in position on the drilling machine table for drill operation to be carried out.

Method of location The workpiece is located by using a cavity of same profile


as that of the workpiece. To prevent the interference between the workpiece edge
and the cavity edge an undercut is provided on the cavity edge corners (refer to
Fig. 6.95).
Locating and Clamping Methods 485

Fig. 6.99 Screw clamp (a) Solid model (b) Detailed drawing

Fig. 6.100 Clamp bush (a) Solid model (b) Detailed drawing
486 Tool Design

Fig. 6.101 Clamp bush liner (a) Solid model (b) Detailed drawing

Fig. 6.102 Hex nut (a) Solid model (b) Detailed drawing

Method of clamping After placing the workpiece inside the cavity on the bush
plate of the drill jig, four screw clamp are used for clamping the workpiece. The loos-
ened clamps can be rotated inside the liner and brought to required position before
Locating and Clamping Methods 487

tightening of the hex nuts. Figure 6.103 shows that three screw clamps are tightened
in required position, whereas, the fourth is loose. For removal of the workpiece all the
hex nuts has to be loosened and the screw clamps rotated free from the workpiece
surface.

Fig. 6.103 Solid model of the designed drill jig (a) Showing the clamping side (b) Showing
the drilling side

Fig. 6.104 Exploded view of the designed drill jig


488 Tool Design

Summary

Locating and clamping element forms are an integral part of jigs and fixture. The
accuracy of the job depends on accurate location and rigid clamping. This chapter
deals with principles, method and devices used for location and clamping.
• The term location refers to the method of establishing correct relative position
of the workpiece with respect to the cutting tool.
• For locating external plain surface, 3-2-1 principle of location is generally
employed.
• For locating cylindrical or circular surface, it is necessary to locate the axis of
the circular workpiece.
• For locating irregular shape workpiece, the profile of the workpiece can be
located by confining the profile of the workpiece by cylindrical locators.
• Adjustable location pins are suitable for locating a rough, uneven or tapered
plane surface.
• The technique used for press-holding the workpiece against the locating sur-
face is defined as clamping and the mechanism used for this is called clamps.
• Clamping forces should be directed toward the nest or locating surfaces.
• The clamping method should be capable of holding the workpiece securely
against the forces developed during operation. However, it should done so
without damaging the workpiece with excessive pressure.
• The position of the clamps should be such that cutting forces should act against
solid part of jig./fixture and not against the clamps.
• Clamps should be quick acting and as simple as possible.
• The clamping points should be provided with ample radius to make the clamp
operable even if there is variation in the workpiece.

Questions
1. What is meant by locating?
2. What is the major factor that determines how a workpiece will be located?
3. How is the configuration of a typical workpiece determined?
4. What is an irregular surface?
5. What is generally used as the basic reference plane for locating?
6. What is meant by complete location?
7. What is necessary to establish complete location of a workpiece whose con-
figuration is formed by flat planes?
8. Why should locating points be placed as far apart as possible?
9. What is the basic reference for circular surfaces?
10. What are the common methods of locating from circular surfaces?
11. What is the best general V angle for locating circular surfaces?
12. What is one of the shortcomings of V location?
13. What is meant by concentric location? Give an example.
14. How does the degree of roughness determine whether a surface is fiat, circular
or irregular?
Locating and Clamping Methods 489

15. How is unevenness compensated for when locating against an irregular sur-
face with more than three locating points?
16. When is sight location used, and how is it accomplished?
17. Why should the tool designer select locating methods that require standard,
readily purchasable locating devices?
18. What are diamond pins, and how are they used?
19. What is the difference between rest pads and rest plates? How are they
used?
20. Why are the bearing surfaces of rest plates sometimes grooved?
21. What is meant by the term nest?
22. How are inaccuracies caused by chips interfering with locating points
controlled?
23. What are the four essential requirements of clamps and clamping devices?
24. What are the basic rules for applying clamping forces?
25. What is a spherical washer, and when is it used?
26. What is the major limitation of cam clamps?
27. What is the limitation and the advantage of latch clamps?
28. What is a quarter-turn screw?
29. How is it possible to use line pressure (form air lines installed in manufacturing
firms) to obtain adequate pressures to cylinders for clamping purposes?
30. Besides being fast-acting and positive, hydraulic and pneumatic clamping
devices have what other advantages?

Problems
1. Figure 1.105a shows a locating plug used to locate a round flanged workpiece.
Can the flanged workpiece be easily removed or will it have a tendency to
stick?
2. An air-to-hydraulic booster unit is installed to operate clamps on a work-hold-
ing fixture. It is connected to the plant air line, which holds a constant 0.6 MPa
pressure. What pressure output can be expected if the ratio of the booster is
16.00?

Design Problems

The following design problems are intended to serve as exercises in applying meth-
ods of locating and clamping. A freehand sketch is to be made of each workpiece in
Fig. 6.105. The sketch should clearly show the method of location and clamping the
workpiece in position as directed in each of the following problems.
1. Locate and clamp the workpiece shown in Fig. 6.105b. The workpiece should
be positioned to allow the slot to be machined on a horizontal milling machine
using a staggered-tooth side-milling cutter.
2. Locate and clamp the workpiece shown in Fig. 6.105c. Position the workpiece
to allow the holes to be drilled on a standard drill press.
3. Locate and clamp the workpiece shown in Fig. 6.105c using the holes as locat-
ing points.
490 Tool Design

Fig. 6.105 Workpieces


Locating and Clamping Methods 491

4. Locate and clamp the workpiece shown in Fig. 6.105d. Position the workpiece
to allow the hole to be drilled on a standard drill press.
5. Locate and clamp the workpiece shown in Fig. 6.105d. Position the workpiece
to allow the 20 mm. slot to be milled on a vertical milling machine using a
20 mm. end mill as a cutter.
6. Locate and clamp the workpiece shown in Fig. 6.105e. Position the workpiece
so that both holes can be drilled in line on a standard drill press.
7. Locate and clamp the workpiece shown in Fig. 6.105e. Locate and clamp in a
lathe so that the 20 mm. -diam shank can be turned with a single-point tool.
8. Locate and clamp the workpiece shown in Fig. 6.105f so that the dovetail slot
can be milled on a vertical milling machine. Note that this workpiece is a rough
casting.

References
Books
• ASTME: “Handbook of Fixture Designs,” F W Wilson (ed.), McGraw-Hill, New York, 1962.
—: “Tool Engineers Handbook,” 2d ed., F W Wilson (ed.), McGraw-Hill, New York, 1959.
• Colvin, F H, and Haas, L L: “Jig and Fixtures,” 5th ed., McGraw-Hill, New York, 1948.
• Grant, H E: “Jigs and Fixtures,” McGraw-Hill, New York, 1967.
• NYSVPAA: “Jig and Fixture Design,” vol. 1, Delmar Publishers, Albany, NY, 1947.
Catalogues
• Accurate Bushing Company, Garwood, NJ.
• American Drill Bushing Company, Los Angeles.
• Carr lane Manufacturing Company, St. Louis, Miss.
• Jergens, Inc., Cliveland, Ohio.
• Universal Engineering Company, Frankenmuth, Mich.
• Vlier Engineering Corporation, Los Angeles.
7 DESIGN OF DRILL JIGS

7.1 INTRODUCTION
Throughout the tool designer’s career he will be concerned with holes. There is hardly
a product produced in American industry that does not contain one or more holes. The
location finish, and size of these holes, may be critical, as in the case of a component
for a missile, or they may just be holes, like those punched in a template for the pur-
pose of hanging it on the wall when it is not in use.
Holes are produced in a variety of ways. They are drilled, reamed, bored, punched
ground, sawed, flame-cut, etc. Drilling is by far the most common method. It has been
estimated that between 80 and 90 per cent of all the pounds of chips made by metal-
cutting operations in the United Sates are produced by drilling. Therefore, a large
portion of tool designer’s time will be spent in designing tools to properly locate drilled
holes. Some tools are simple and easily designed. Others are complex and need
careful development. The purpose of this chapter is to introduce the prospective tool
designer to the basic types of tooling for drilled holes, although reaming, tapping, and
boring will also be discussed.

7.2 DEFINITION OF A DRILL JIG


A drill jig is a device for ensuring that a hole to be drilled, tapped, or reamed in a work-
piece will be machined in the proper place. Thus, instead of laying out the position
of each hole on each workpiece with the aid of a square, straightedge, scriber, and
centre punch, the operator uses a jig to guide the drill into the proper place. Basically,
it consists of a clamping device to hold the part in position under hardened-steel bush-
ings through which the drill passes during the drilling operation. The drill is guided
by the bushings. If the workpiece is of simple construction, the jig may be clamped
on the workpiece. In most cases, however, the workpiece is held by the jig, and the
jig arranged so that the workpiece can be quickly inserted and as quickly removed
after the machining operation is performed. Figure 7.1 shows an adjustable drill jig
designed for precision cross-hole drilling of round or hexagon stock.

Fig. 7.1 Adjustable drill jig (Heinrich Tools, Inc.)


Design of Drill Jigs 493

Jig make it possible to drill, ream, and tap holes at much greater speeds and with
greater accuracy than the holes are produced by produced by conventional hand
methods. Another advantage is that skilled workers are not required when jigs are
used. Responsibility for the accuracy of hole location is taken from the operator and
given to the jig. The hole pattern should be within the required tolerances on each part.
This is the basis for the interchangeable-part concept that has made Eli Whitney’s
name so prominent in industrial history.
The term jig should be used only for devices employed while drilling, reaming, or
tapping holes. It is not fastened to the machine on which it used and may be moved
around on the table of the drilling machine to bring each bushing directly under the
drill. Jigs physically limit and control the path of the cutting tool.
If the operation includes machining operations like milling, planing, shaping, turn-
ing, etc., the term fixture should be used. A fixture holds the work during machining
operations but does not contain special arrangements for guiding the cutting tool, as
drill jigs do. A fixture is also fixed to the machine. Fixtures will be discussed in detail
in Chap. 8.

7.3 TYPES OF DRILL JIGS


There are so many types of drill jigs it is difficult to classify them into a standard sys-
tem. Various companies and industries may refer to the same type of jig by different
names. There are types of jigs which have become common to most industries, and
with exceptions these types may be considered as ‘standard’ by general agreement.
They are leaf, box, tumble, template, plate, indexing, universal, and vise. Several other
types are in existence, as the type of jig is limited only by the imagination of the tool
designer. However, example, of the above common types will introduce the prospec-
tive tool designer to the subject and prepare him for an understanding of the principal
parts.

Leaf jigs This type is distinguished by its hinged cover, or leaf, which is swung open
to load or unload the jig (see Fig. 7.2). After the workpiece has been located inside the
jig, the leaf is firmly closed and locked.

Locking screw
Quarter-turn
Hinge
Leaf (hinged plate) screw
Side-locating
Bushing Fixed stop
stop
Slot

Bearing Chip End-locating Jig feet


surface clearance stop

Fig. 7.2 Simple leaf jig


494 Tool Design

Leaf jigs can be loaded and unloaded quickly and are suitable for complicated work-
piece with irregular contours. Holes may be drilled in more than one surface for a
single loading of the jig. Drill bushings may be located in the leaf and reamer bushings
in the base, providing the leaf is rigid enough to withstand the pressure exerted by the
reamer (see Fig. 7.3).

Drill bushings

Reamer bushings Workpiece

Fig. 7.3 Leaf jig utilising drill bushings in hinged plate and reamer bushings is base

The disadvantage of leaf jigs is that chips may accumulate inside and cause trouble
unless provisions are made for them. Also if drill bushings are fitted in the leaf, play in
the hinges may affect drilling accuracy. For this reason, it is the best practise to locate
drill bushings in fixed portions of the jig whenever possible, as shown in Fig. 7.4. Often
this is not possible, and the bushing must be located in the leaf or hinge plate. When
this is done, the hinge must be precision-made to ensure accurate alignment and the
joints should be constructed of heat-treated pins and bushings to prevent excessive
wear. When bushings are located in the hinge plate, the hole pattern is fixed, although
the hinge plate is movable. The only chance for error is in the location of holes relative
to the part outline which may or may not be critical.

Fig. 7.4 Leaf-type diameter jig showing drill bushings located in main body

It is considered poor design to clamp the part with the hinged plate. The hinged plate
should close against a fixed stop, as shown in Fig. 7.2. There are times, however,
when the hinged plate serves as a support when reaming. Note that the thumbscrew
acts as a clamp after the hinged plate is locked in place.
Design of Drill Jigs 495

Box and tumble jigs This type receives its name from its shape which, in general,
resembles a box. Figures 7.5 and 7.6 are typical examples. In Fig. 7.5, the workpiece
is a steel block used as a specimen in an oil-testing machine. The hole is drilling and
then reamed while in the jig. The loading is as follows: the cam rod is taken out of the
jig and the workpiece placed in position inside the jig. The cam rod is then replaced
and rotated to its locking position. This holds the workpiece firmly so that the drilling
operation can be performed.

Drill bushing Handle


Cam rod
Box Bushing

Workpiece

Cam

Fig. 7.5 Box jig

Fig. 7.6 Box jig utilising hydraulic clamping methods (Vlier Engineering Corp.)

When a box jig contains bushings on two or more sides for the purpose of drilling
holes on different sides of the part, it is referred to as a tumble jig. Such a jig has sets
of jig feet on opposite sides of the work faces. After one face is drilled, the next side
may be drilled by simply flopping the jig to expose this side to the drill spindle. The
advantage of tumble jigs is that greater accuracy can be obtained and less part han-
dling is necessary Figure 7.7 shows an example of a tumble jig.
Sometimes, box jigs are constructed in a channel-shaped trough and clamped by
means of a screw, as shown in Fig. 7.8. This type is limited to jobs having workpiece
of simple symmetrical shape. They are often referred to as channel jigs.
496 Tool Design

Fig. 7.7 Box-type tumble jig

Bushing

Knob screw

Workpiece

Fig. 7.8 Channel jig

Template jigs This is a simple type used for locating hole patterns on large work-
piece. It may contain locating devices, or it may be necessary for the operator to
locate by means of measurement. Some have provisions for clamping. It should be
emphasised that although the relationship of a hole pattern to the part outline is not
accurate, the hole pattern itself is quite accurate. Figure 7.9 shows an example of a
template jig.

Plate jigs The plate jig is so named because a plate containing bushings forms the
main structural member. It is of open construction, which facilitates part loading and
removal, chip removal and clamping. For this reason, it is sometimes referred to as an
open jig. Slip bushings may be used with liner bushings to allow operations other than
drilling. Thin parts may be stacked for drilling several parts at time. The jig may or may
not have legs. It is generally inexpensive to construct since it is largely fabricated from
standard parts and stock-size plates.
Design of Drill Jigs 497

Fig. 7.9 Template jig

Figure 7.10 shows a typical plate jig. The loading and unloading is accomplished by
turning the jig over from the position shown in the drawing. The jig then rests on the
small end of the legs in A. The workpiece, a carburettor part, is placed over the main
bolt and located with the aid of two gauge plates. The U-shaped washer is then put in
place as shown and the nut tightened. This locks the workpiece firmly. The jig is then
turned over to the drilling position, and the two holes are drilled in the flange of the
workpiece. Note that the main bolt is cut away to shoulder in B so that it is possible to
load and unload the workpiece without removing the nut.

Fig. 7.10 Plate jig


498 Tool Design

The disadvantage of plate jigs are that only one surface can be drilled at one loading
and drilling forces are generally directed toward the clamping device. It is therefore
necessary that the clamping device be rigid enough to withstand drilling forces.

Indexing jigs Indexing jigs are used for circular hole patterns in which the part is
indexed successively to the different positions under a single bushing. The location of
the holes may be taken from the first hole drilled, as shown in Fig. 7.11, or from other
holes in the part. Sometimes and indexing plate or device is incorporated in the jig,
as shown in Fig. 7.12.

Fig. 7.11 Simple indexing jig designed for drilling four holes in steel collar

Fig. 7.12 Indexing jig made from standard indexing fixture

This jig was developed from a basic indexing fixture with the workpiece held in a
chuck as shown. The single drill bushing can be adjusted to different heights, as well
as laterally, to allow holes to be drilled on the pitch circle of different sizes. It will readily
be seen that this jig has several universal features and thus could be called a univer-
sal jig, but notwithstanding this, the classification index jig is more descriptive.
Design of Drill Jigs 499

Universal or pump jigs These jigs are pro-


duced commercially as basic units and are
adapted to specific jobs by tool designers and
tool makers. Many fine types are available and
the outstanding features of these jigs are their
rigidity, low height, ample chip clearance, and
ease of operation. The moving parts are com-
pletely protected from chips. The working parts
consist of a handle connected to a cam or rack
which moves either a bushing plate or a nest
vertically in order to clamp a workpiece (see
Fig. 7.13).
Some pump jigs incorporate a swing-away fea- Fig. 7.13 Universal or pump jig
ture, which permits faster loading easier nest- (The Do All Company)
ing, and availability of the part for inspection (see Fig. 7.14).

Fig. 7.14 Pump jig with swing-away feature (Accurate Bushing Company)

Vise jigs Vise jigs are constructed by attaching special inserts to the jaws of a regu-
lar machine vise. The inserts can be made quickly with a minimum amount of expense.
They are usually used for low volume and short production runs but may be used for
long production runs of simple parts. Figure 7.15 shows this type of application.
500 Tool Design

Fig. 7.15 Typical vise jigs (Heinrich Tools, Inc.)

7.4 CHIP FORMATION IN DRILLING


Before entering into further discussion on the design of drill jigs, it would be better
to review the mechanics and chip formation of drilling. The tool designer must be
acquainted with the process to design the most effective tools for use with a drilling
operation.
As a two-flute drill passes through a piece of metal is displaced by two parts of the drill
point, the drill lips and the chisel edge of the drill. The cutting process of the drill lips is
essentially the same as that of a lathe cutting tool, but at the chisel edge, the process
is quite different. Here the deformation of the metal is much more severe as the chisel
edge tends to penetrate the metal and extrude it outward until it is picked up by the
drill lips. The nature of the cutting action at the chisel edge is the reason twist drills
require such large axial thrust forces. These drilling forces can be reduced by proper
drill grinding, thinning the drill web, and grinding special drill points such as crankshaft
points spiral points, and prismatic points (see Chapter 4).
After the metal is displaced by the chisel edge and the cutting lips of the drill, it must
be ejected from the hole. The drill flutes act as an auger to remove the chips, and
since there is little room in the flutes, it is desirable to have the chips broken up into
small pieces. Long coils tend to pack in the flutes and retard the flow of coolant to the
cutting edges. The result is excessive heat and premature drill dulling.
The factors that affect the breaking or coiling of the chip are the ductility of the mate-
rial being drilled and the thickness of the chip. The chips tend to bend and curl when
drilling the more ductile materials. The same is true when the chip is thin. In less
ductile materials, the chips tend to break into smaller pieces, which is the desirable
condition.
It is sometimes necessary to alter the drill geometry in order to break the chips. This
can be done by decreasing the rake angles of the cutting lips, which causes the chips
to bend more sharply and break. Special drills with built-in chip breakers or ground-in
chip breakers may also be used.
Design of Drill Jigs 501

Chip coiling can sometimes be remedied by


increasing the feed, which results in increasing Minor burr
the thickness of the chip. Using fast spiral drills
will also increase the ability of a drill to remove
chips from the hole.
Major burr
Burrs are invariably raised on the work during a
drilling operation, as shown in Fig. 7.16. The burr Fig. 7.16 Major and minor burrs
raised at the start of the hole is usually smaller
than that at the end of the hole and is referred to as a minor burr. The burr at the end
of the hole is formed as the drill breaks through the material and is referred to us a
major burr. Major and minor burrs must be taken into consideration when designing
drill jigs since they may cause difficulty in removing the part from the jig after the hole
is drilled.

7.5 GENERAL CONSIDERATIONS IN THE DESIGN


OF DRILL JIGS
Part of the following information has been discussed in Chapter 6. Any repetition at
this point will have to do with drill jigs. The purpose of this section is to point out factors
of jig design that influence the results during drilling operations.

Rigidity Rigidity is first and foremost in the design of any tooling. It is far better for
a tool to be built too rigid than not rigid enough. A drill jig may produce faulty work if
it bends only a few microns. It must be strong enough to withstand all forces applied
to it and must also resist deflections which would destroy its accuracy. Such defec-
tions may be the result of excessive tightening of clamps that hold the workpiece in
place. The work must also be supported so that it does not bend under drilling pres-
sure. Such bending may result in holes that are oversize and out of line and may be
the cause of excessive drill breakage as the drill may grab when the drilling pressure
is released at the end of the hole. Grinding split or crankshaft points on the drill will
reduce drilling pressure and may avoid excessive deflection of the work. However,
extra tool grinding will increase the cost of the workpiece.
It should be noted that drill jigs often receive rough handling and therefore must be
strong enough to resist this type of treatment. Small projections which can catch on
the storage racks and be bent should be beefed up, eliminated, or designed into the
main body of the jig in such a way that they will not catch.
The tool designer usually depends upon his past experience when determining the
size of the component parts of a drill jig. He has a good idea of how thick the bush-
ing plate must be to withstand the forces it will be subjected to during the drilling
operation. This knowledge comes only by experience, and the beginning tool designer
would do well to study examples of proved designs before he begins his own.
This discussion is not to imply that a stress analysis is never necessary in the design
of drill jigs. Larger jigs may require a stress study since it may lead to a great deal of
weight reduction. Small jigs usually do not require such a study.

Location and clamping The pocket or resting position of the workpiece in a jig
or fixture is referred to as a nest. The jig should be designed in such a way that the
502 Tool Design

workpiece can be fitted easily into the nest by the operator. Enough clearance should
be provided to allow the part to slip in and out of the nest without difficulty. This should
be possible with only moderate care on the part of the operator. Clamping methods
should be rapid and positive, and the clamps should be located in a position for the
greatest convenience of the operator. Clamps should be located over (or as near as
possible to) a bearing point on the workpiece to prevent distortion of the part and
should be positioned to resist the pressure of the cutting tool. They should hold the
part firmly in the nest, and the design should be such that the drill will tend to push
the part against the nest and not against the clamp. Clamps should not extend exces-
sively above the drill bushings, as they may strike protruding tools when moved from
one tool to another.
Complicated clamping devices should be avoided. Many unusual devices appear in
various handbooks and in Chapter 6 of this text, but they are expensive to build and
may be a source of trouble. Clamping devices should have a minimum number of
loose or moving parts. It is better to use standard commercial clamping units when-
ever possible. The tool designer’s nightmare type of clamp with moving parts, com-
pound levers, and around-the-corner devices, is fine to fill the pages of a handbook
but has no place in the large majority of drill-jig applications. It requires a great deal of
maintenance which may offset the savings of faster operation. A simple clamp with a
minimum amount of parts will serve in most cases.
Loose parts such as handles, pins, wrenches, gauges, etc., are another source of
trouble in the production shop. They will be lost in spite of all precautions taken by
production personnel. Whenever possible, a handle of a clamping device should be an
integral part of the device. If this is not possible, the handle should be fastened to the
jig by a small chain or cable. Suitable cable assemblies like those shown in Fig. 7.17
are available commercially for such applications. If the loose part is such that it cannot
be fastened to the jig body, it should be marked with proper identification, and some
provision for storage should be provided on the body of the jig.

Fig. 7.17 Using nylon cable assemblies to fasten loose parts to jig body (Carr Lane Manu-
facturing Co.)

The locating points on a drill jig should be designed in such a way that it is impossible
to load the workpiece incorrectly. They should be placed as far apart as the shape of
Design of Drill Jigs 503

the workpiece will allow and be of a design that will allow easy cleaning and minimum
chip build-up. Locating points should be in view, if possible. Locators subjected to
wear should be easily replaced and made of hardened tool steel. Figure 7.18 shows
a spherical radius locator in common use.
Work-supporting or bearing surfaces are provided in the form of pins, pads, or plates.
Pads and plates may be machined into the base or nest of the jig or purchased as
standard parts, as shown in Fig. 7.18. Standard jig feet are sometimes used as sup-
porting surfaces. One advantage in the use of standard pins, pads, and plates is that
they are heat-treated and therefore less subject to wear. They can also be replaced
when excessive wear occurs.

Fig. 7.18 Standard rest pads, buttons, plates, and locators

Chip control For all practical purposes, two basic types of chip will be encoun-
tered in drilling, segmental and continuous. Segmental chips are in the form of small
particles or segments, a typical example being chips produced in drilling cast iron.
Continuous chips are long and stringy and tend to take the form of long coils or spirals.
Such a chip may be produced when drilling ductile steel.
The segmental type of chip is preferred when drilling because it is more easily con-
trolled and removed from the drill jig. Continuous chips can be made to break by
decreasing the rake angle of the drill, using a drill with a slower spiral, grinding in chip
breakers, of using mechanical oscillator of the feed of the drill press to provide an
intermittent feed (see Chapter 4).
In the design of drill jigs, enough space should be provided between the work and the
bottom of the drill bushing to allow chips to pass between the work and bushing plate
rather than through the drill bushing (see Fig. 7.19). This helps eliminate chips packing
in the drill flutes, allows more coolant to reach the cutting edge of the drill, and helps
to prevent wear of the abrasive chips on the drill bushing. This is an ideal situation and
works well for discontinuous chips. The beginning tool designer will learn, however,
that the ideal situation in quite often not the case. If the chips are not segmental, they
may tangle between the work and the bushing and be difficult to remove (see Fig.
7.20). In this case, it may be better to locate the bushing closer to the work, which will
force the long stringy chip through the bushing where any entanglement can easily be
removed by the operator.
504 Tool Design

Bushing Bushing plate

Segmental chips

Work

Fig. 7.19 Chips do not pass through bushing

In any case, the drill bushing should not be placed so far away from the work that the
drill is allowed to bend. The general rule is to place bushings at a distance of 1 to 1.5
times the drill diameter from the workpiece. Zero clearance between the bushing and
the work will give the maximum guiding effect, and for angular entry, as shown in Fig.
7.21, it is best to have zero clearance. Hence, again the chips must pass through the
bushing and cause excessive wear. Also a deep hole effect occurs because the drill
creates a hole equal to its depth. It is more difficult for chips to come up through the
drill bushing, but this may be partially eliminated by using fast spiral drills.

Bushing
Bushing
plate

Work

Fig. 7.20 Chip tangle between bushing plate Fig. 7.21 Drill-bushing position for
and workpiece angular drill entry

Chips are generally removed from jigs in three ways. Coolant will help to flush away
chips as they are formed, they may be removed manually with a brush or hook, or
they may be removed with compressed air. Coolant paths should be provided to aid
chip removal by coolant. If chips are to be removed manually with a brush, corners
and obstructions should be eliminated to allow the operator to brush out chips with a
minimum amount of effort. When compressed air is used, guards should be provided
to prevent scattering of chips throughout the working area.
A word of caution about the use of compressed air as a chip-removal medium. Small
chips and dirt will be blown into bearing surfaces and cause wear and loss of accuracy.
Many companies do not allow compressed air to be used around expensive machine
tools. Companies air should be used with common sense. However, many operators
lack common sense when it comes to using compressed air around machine tools
and will direct the air blast into bearings and bearing surface. Unless the saving in
time by using compressed air will offset the cost of the machine and the loss of accu-
racy, chips should be removed by other means.
Design of Drill Jigs 505

Small chips have a tendency to build up on the corners of nests or locators and
prevent the workpiece from seating properly (see Fig. 7.22). Provisions should be
made in the design of drill jigs for corner relief, which will help prevent chip buildup in
corners. Figure 7.23 shows various means of providing chip relief.

Locators A

Work

Sec. A-A
A

Fig. 7.22 Improper design using block type locators. Chips cannot escape from corners.

Fig. 7.23 Methods of providing corner relief: (a) Using button locators to provide corner
relief and (b) Block locations with milled-in corner relief

Small chips may also stick to contact surfaces and cause the part to seat improperly. If
contact surfaces are reduced in area, this tendency may be reduced sine the area on
which chips can be gather is minimised. This is known as providing area relief. Figure
7.24 shows an example of area relief. The area under the part must not be reduced to
the point where proper support is no longer provided.
506 Tool Design

Dowel pin, typ. Socket-head cap


Area relief screw, typ. Corner relief

Sight
holes
Locating points

Support under
locking device
Locating
point

Jig feet on Drill


extremities of jig bushing, typ.

Open area for


chip passage Note: Sight holes
enable operator to
check locating points
Backup hole slightly
larger than hole in
part to allow for major
burr
Area relief

Fig. 7.24 Built-up drill jig showing area relief, corner relief, location of standard jig feet,
and locating points

Jig feet and legs A drill jig should stand on four feet (or legs) rather than a flat sur-
face. Chips may get under a flat surface and tip the jig. Four legs are recommended
as three will not tip if chips get under one leg. The operator may not be aware that the
jig is tipped if three legs are used.
Jig feet may be built into the jig body or purchased as standard parts. Types of stan-
dard jig feet are shown in Fig. 7.25. They should be placed on the jig so that all bush-
ings are within the area bounded by the legs, which usually are on the extremities of
the jig(see Fig. 7.24). Feet should be ground so that they are all in one plane after they
are mounted on the jig base.

Miscellaneous considerations Provisions should be made for the coolant to get to


the drill. For example, coolant may have to come in from the side to be applied directly
to the drill, rather than coming down through the drill bushing. There should also be
holes or passages for escape of the coolant. The operator should not have to tip the
jig to drain excessive coolant from nests or pockets in the drill jig.
Thought should be given to operator safety when designing a drill jig. Sharp corners
and projections should be eliminated. The jig should be large enough to enable the
operator to hold the jig against the torque of the drilling machine. It may be necessary
to increase the operator’s advantage by adding a handle to the jig when the opera-
tor’s hands are covered with coolant. The jig should be as light as possible, since the
operator must handle it constantly.
Design of Drill Jigs 507

Rest button Hexagonal rest button Jig leg

Cap-screw rest button Flat jig feed

Fig. 7.25 Types of standard jig feet

7.6 DRILL BUSHINGS


Bushings are used to guide drills, reamers, and other cutting tools into the proper
position on the workpiece. They are of tool steel and are hardened to RC60 to 64 to
provide a wear resisting surface. They are precision-made, the outside being ground
and the inside either ground or lapped to within 0.007 mm concentricity.
The length of the bushing is quite important. Generally speaking, the length of a bush-
ing should be approximately twice the diameter of the bushing hole. The diameter of
the bushing hole should be very close to the diameter of the drill but should not be so
tight that the drill will drag in the bushing. A general rule for clearance between the
drill and the bushing wall is from 0.01 to 0.025 mm. Too great a clearance may cause
chipped drill margins and hole inaccuracy.
Fortunately, the tool designer does not have to be concerned with the design of drill
bushing except in special cases. Standard dimensions have been adopted by the
American National Standards Institute, and with few exceptions, all bushing manufac-
turers have adopted these standards. It is only necessary for the tool designer to refer
to manufacturers’ catalogues for bushing sizes. Special sizes may be obtained from
many manufacturers upon request. Special designs may have to be manufactured
by the tool room within the tool designer’s own company or by local tool shops that
specialise in this kind of work.
ANSI has classified jig bushings as (i) press-fit bushings, (ii) renewable bushings,
and (iii) liner bushings. Press-fit (wearing) bushings are for direct installation into the
bushing plate without the use of a liner and are used for short-run production where
bushings do not require replacement. They also may be used where the closeness
of the centres of holes will not allow the installation of liners and renewable bushings.
Press-fit bushings are made with or without heads, as shown in Fig. 7.26.
508 Tool Design

(a) (b)

Fig. 7.26 Press-fit wearing bushing (a) Type P headless press-fit bushing, used for permanent
press-fit installations, preferably under light axial loads; ideal where top of bushing
must be flush with jig plate or where dimensions between bushing are too close
for bushing heads (b) Types H head press-fit bushing, used for permanent press-
fit installations; can be mounted flush with jig plate by counter boring mounting
hole to fit bushing head. Head resists effect of heavy axial loads. (American Drill
Bushing Company)

Renewable bushings (wearing) are for use in bushing liners that have been installed
in the bushing plate. They are divided into two classes, fixed and slip (see Fig. 7.27).
Fixed renewable bushing are installed in a liner with the intention of leaving them in
place until they wear out. Slip renewable bushing are made with a knurled head that is
machined for a locking device. Slip renewable bushings are made with a knurled head
that is machined for a locking device. They are used when two or more operations
such as reaming, tapping, spot facing etc., are performed with the same jig.

(a) (b)

Fig. 7.27 Renewable wearing bushings (a) Type S slip renewable bushing, used with liners
in applications requiring frequent change of bushing to perform mote than one
operation in a hole, such as drilling and reaming. Body OD is ground to provide
correct slip fit with liner. Head is designed to permit quick removal or bushing
without removing lock screw. (b) Type F fixed renewable bushing, used with lin-
ers in long production runs. For single operations such as drilling only. Quickly
replaced when worn out by removing lock screw and inserting new bushing. Body
OD is ground to provide correct slip fit with liner. Head has milled lock-screw
recess for securing bushing in place. (American Drill Bushing Company)
Design of Drill Jigs 509

Bushing liners, sometimes called master bush-


ings, are supplied with or without heads. The
special liner shown in Fig. 7.28 eliminates the
need for a locking device.
Special bushing are sometimes necessary to
meet the requirements of the workpiece, For
example, two or more holes may be so close
together that it is impossible to have an individ-
ual standard bushing for each hole. In this case,
standard thin-wall bushings may be used. Thin- Fig. 7.28 Typical bushing liner
wall bushing are basically the same in tolerance (American Drill Bushing
and application as ANSI standard bushings, the Company)
prime difference being in the thinner wall and the
reduced head dimensions.
Another method of overcoming close hole distance is to use two holes in one standard
bushing, as shown in Fig. 7.29. Or one hole may be placed in an eccentric bushing,
as shown in Fig. 7.30. The bushing is indexed 180° for drilling the second hole. Two
standard bushings are sometimes modified for close hole drilling (see Fig. 7.31). A
flat is ground on the body of each bushing. The bushings are pressed into the bushing
plate and are prevented from turning by the interlocking of the two flats.

Fig. 7.29 Use of two holes in one bushing Fig. 7.30 Use of eccentric for close
hole drilling

Fig. 7.31 Modification of standard bushings for close hole drilling


510 Tool Design

Screw bushings, like those shown in Fig. 7.32, are sometimes used where it is nec-
essary for the bushing to hold the workpiece as well as guide the drill. For the most
part, this a poor practise because of the difficulty of maintaining accuracy in the screw
bushing. This is overcome by marking the thread loose and guiding the bushing by a
ground cylindrical portion, as shown in Fig. 7.33. It is apparent that this type of bush-
ing construction would be expensive to machine accurately.

Fig. 7.32 Plain screw bushing Fig. 7.33 Screw bushing not guided
by thread

Drill-bushing manufacturers are presently marking bushings for embedment in cast


able nonferrous metals, plastics, or wood. Such bushings are shown in Fig. 7.34. The
use of these bushings is discussed further in Chapter 11.

Bushings as guides for reamers Reaming in jigs is accomplished in the same way
as drilling. Perhaps the only difference is that the operator must exercise more care in
aligning the bushing with the reamer. A reamer is by nature a more fragile tool than a
drill, and if the reamer is forced to drag the jig over into alignment, the result may be
excessive wear and chipped cutting edges of the reamer. This may be partially over-
come by use of a pilot section on the end of the reamer.
The accuracy of reaming may be improved by use of guides on both sides of the
workpiece, especially if the hole is quite deep. In Fig. 7.35, a reamer with two pilots is
used. Note that the pilots are grooved to permit cutting oil to enter the bushings and
aid in lubrication. The grooves also act as a chip cleaner.

(a) (b)

Fig. 7.34 Bushing for embedment


(American Drill Bushing Fig. 7.35 Use of double guide bushings
Company) in a reaming jig
Design of Drill Jigs 511

7.7 METHODS OF CONSTRUCTION


The most common method of constructing drill jigs is by holding the components
together mechanically, often referred to as built-up construction. The advantage of this
type of construction is that all the parts can be completely machined before assembly
and worn parts can easily be replaced. Minor adjustments can also be made during
assembly.
Jigs of this type are held together by socket-head cap screws and dowels. The screws
serve only to hold the components together, while the dowels serve to hold the parts
in alignment. A minimum of one screw and two dowels must be used to fasten each
component to the jig body. More screws may be necessary, but two dowels are usually
sufficient to hold the components in alignment. Dowel holes should be drilled com-
pletely through both the component and jig body to facilitate easy removal. Dowels
pressed into a blind hole are extremely hard to remove. If it is necessary to use a blind
hole, it should be reamed 0.01 mm larger than the other hole. This will make the dowel
stay in the through hole when the assembly is taken apart.
Dowels should be located as far apart as possible to give the best alignment. They
should be pressed into both jig components. Toolmarks usually employ a set of dowel-
pin reamers which are ground undersize when preparing dowels holes.
Welded construction is often used in building a drill jig. In certain designs, it may
be faster to weld the components together and then machine them to finished size.
Welded jigs must be stress-relieved before machining because the relief of stresses
set up during welding will warp the jig.
Drill jigs are sometimes constructed by a combination of welding mechanical assembly.
The main body may be of welded construction with the smaller components attached
by screws and dowels. In this case, parts subjected to wear can easily be replaced.
The main body of a drill jig may also be cast, but this is expensive if only one jig is
being built. When a standard cast body can be used for several jigs, cast construction
certainly has merit.
Cast construction provides a neater jig since the metal may be distributed to better
advantage. It may also be easier to cast odd shapes than it is to machine them.

7.8 DRILL JIGS AND MODERN MANUFACTURING


Drill jigs have been in use since the beginning of modern manufacturing processes.
Their main advantages are the reduction of labour cost and machining operations,
inter. changeability of parts, and the possibility of supplying replacement parts. Drill
jigs will continue to be used in modern industry, especially when a large number of
parts are to be produced.
There is, however, a problem when a limited number of workpieces are to be pro-
duced. The small number of workpieces may not warrant the cost of a jig. This has
led to the use of jig-boring machines and precision boring machines as production
tools. Although they were originally designed as tool room machines, their ability to
permit the accurate location and boring of holes has made them well suited for small-
lot production without the use of drill jigs. The disadvantage in this type of production
512 Tool Design

is that the drilling or boring operations are slow and a relatively skilled machinist is
required.
The addition of a positioning control system to precision drilling and boring machines
greatly decrease production time. An example would be a numerically controlled drill
press. Once a tape has been prepared, it is only necessary to mount the workpieces
in a simple fixture and the machine automatically positions and drills the workpieces.
This typs of machine has increased production rates to the points where the design
and construction of drill jigs cannot be justified by previous guidelines. In many cases,
the use of numerically controlled drilling and boring machines has completely elimi-
nated the need for drill jigs.
This is not to imply that drill jigs are on their way out a manufacturing method. However,
the tool planner and designer should investigate the new advances in manufacturing
technology and use them to his advantage.

Solved Design Problems

Problem 1 Design a locating and clamping method to drill all the holes in the work-
piece shown in Fig. 7.36 (a) and (b).

Fig. 7.36 The finished workpiece (a) Solid model (b) Detailed drawing

Solution: The given problem requires designing a drill jig for the workpiece so that
all the eight holes can be drilled. The workpiece can be located u sing a nest or cavity
locator and can be clamped by screw clamping.
The designed parts of the drill jig are presented in the Table 7.1 with reference to the
solid model and detailed drawing of individual parts.
Design of Drill Jigs 513

Table 7.1 Designed drill jig parts


SI. No Part name/Part description Quantity Figure reference

1. Bush Plate 1 7.37 (a) and (b)


2. Drill Bush 8 7.38 (a) and (b)
3. Screw Clamp 3 7.39 (a) and (b)
4. Screw Clamp Bush 3 7.40 (a) and (b)
5. Screw Clamp Liner 3 7.41 (a) and (b)
6. Hex Nuts 3 7.42 (a) and (b)

Fig. 7.37 Bush plate (a) Solid model (b) Detailed drawing

Fig. 7.38 Bush 14 mm (a) Solid model (b) Detailed drawing


514 Tool Design

Fig. 7.39 Screw clamp (a) Solid model (b) Detailed drawing

Fig. 7.40 Clamp bush (a) Solid model (b) Detailed drawing

Fig. 7.41 Clamp bush liner (a) Solid model (b) Detailed drawing
Design of Drill Jigs 515

Fig. 7.42 Hex nut (a) Solid model (b) Detailed drawing

The complete assembly of all the parts is shown in Fig. 7.43 (a) and (b). The exploded
view of the assembly is shown in Fig. 7.44 The drill jig is first tuned over and work-
piece is placed in the cavity and clamped in position. The drill jig is then rotated and
placed in position on the drilling machine table for drill operation to be carried out.
Method of location: The workpiece is located by using a cavity of same profile as
that of the workpiece on the lower side of the bush plate. To prevent the interference
between the workpiece edge. Corners and the cavity edge corners and undercut is
provided on the cavity edge corners (refer to Fig. 7.37).
Method of clamping: The workpiece placed inside the cavity in the bush plate of the
brill jig. The three screw clamps are then used for clamping the workpiece. The loose
clamps can be rotated inside the liner and brought to required position before tighten-
ing of the hex nuts. Figure 7.43 shown that two screw clamps are tightened in required
position, whereas the third is loose. For removal of the workpiece, all the hex nuts has
to be loosened and the screw clamps rotated free form the workpiece surface.

Fig. 7.43 Solid model of the designed drill jig (a) Showing the clamping side (b) Showing
the drilling side
516 Tool Design

Fig. 7.44 Exploded view of the designed drill jig


Problem 2 Design a drill jig to drill a 10 mm hole in the workpiece as shown in
Fig. 7.45 (a) and (b).

Fig. 7.45 The finished workpiece (a) Solid model (b) Detailed drawing

Solution: The given problem requires designing a drill jig for drilling a 10 mm hole on
the workpiece at a distance of 30 mm from the end edge of the workpiece. The axial
hole on the workpiece can be conveniently used for location. The workpiece can be
located using a stud locator and can be clamped using a C washer and knob.
The designed parts of the drill jig are presented in the Table 7.2 with reference to the
solid model and detailed drawing of individual parts.
Design of Drill Jigs 517

Table 7.2 Designed drill jig parts


SI. No Part name/Part description Quantity Figure reference

1. Angle Plate 1 7.46 (a) and (b)


2. Bush Plate 1 7.47 (a) and (b)
3. Drill Bush 1 7.48 (a) and (b)
4. Socket Head Screw 3 7.49 (a) and (b)
5. Locating stud 1 7.50 (a) and (b)
6. Clamping Knob 1 7.51 (a) and (b)
7. C Washer 1 7.52 (a) and (b)
8. Knob (stue rear end) 1 7.53 (a) and (b)

Fig. 7.46 Angle Plate (a) Solid model (b) Detailed drawing

Fig. 7.47 Bush Plate (a) Solid model (b) Detailed drawing
518 Tool Design

Fig. 7.48 Bush (a) Solid model (b) Detailed drawing

Fig. 7.49 Socket head screw (a) Solid model (b) Detailed drawing

Fig. 7.50 Locating stud (a) Solid model (b) Detailed drawing
Design of Drill Jigs 519

Fig. 7.51 Clamping Knob (a) Solid model (b) Detailed drawing

Fig. 7.52 C Washer (a) Solid model (b) Detailed drawing

Fig. 7.53 Knob (stud rear end) (a) Solid model (b) Detailed drawing
The complete assembly of all the parts is shown in Fig. 7.54. The exploded view of
the assembly is shown in Fig. 7.55. The stud is attached to the angle plate using a
520 Tool Design

specially designed knob at the rear end. The bush plate is attached to the angle pate
using three socket head screw. The bush is fitted to the bush plate at the desired
located, so the drilled hole will be located at the correct position.
The workpiece is located by using a stud and the front face of the angle plate. To
prevent the interference between the workpiece edge stud is relieved slightly (refer
to Fig. 7.50). After placing the workpiece on the stud, it is clamped using a C-washer
and a knob.

Fig. 7.54 Solid model of the designed drill jig

Fig. 7.55 Exploded view of the designed drill jig

Problem 3 For the workpiece shown in the previous problem, design a drill jig so
as to drill six equally spaced 10 mm holes of 15 mm depth at a distance 30 mm from
any end of the workpiece.
Design of Drill Jigs 521

Solution: The workpiece is required to be located and clamped, for drilling six equi-
spaced 10 mm holes of 15 mm depth at a distance 30 mm form any end of the work-
piece. This essentially requires designing a indexing drill jig. The locating and clamp-
ing methods remains same as that discussed in the previous problem. However, in
order to drill six hold on the circumference of the workpiece, it requires incorporation
of a indexing mechanism. The indexing mechanism should be capable of dividing the
workpiece in six equal divisions (i.e., 60°). For this purpose, an indexing pin is used
along with an index plate. The index plate is fixed to the angle plate using socket head
screw. The index plate can also be directly welded to the angle plate.
The designed drill is capable of indexing the workpiece only in six equal divisions. The
designed parts of the drill jig are presented in the Table 7.3 with reference to the solid
model and detailed drawing of individual parts.

Fig. 7.56 Angle Plate (a) Solid model (b) Detailed drawing

Table 7.3 Designed drill jig parts


SI. No Part name/Part description Quantity Figure reference

1. Angle Plate 1 7.56 (a) and (b)


2. Bush Plate 1 7.47 (a) and (b)
3. Drill Bush 1 7.48 (a) and (b)
4. Socket Head Screw 4 7.49 (a) and (b)
5. Locating stud 1 7.50 (a) and (b)
6. Clamping Knob 1 7.51 (a) and (b)
7. C Washer 1 7.52 (a) and (b)
8. Knob (stud rear end) 1 7.53 (a) and (b)
9. Index plate 1 7.57 (a) and (b)
10. Indexing pin 1 7.58 (a) and (b)
522 Tool Design

Fig. 7.57 Index plate (a) Solid model (b) Detailed drawing

Fig. 7.58 Indexing pin (a) Solid model (b) Detailed drawing

The complete assembly of all the parts is shown in Fig. 7.59. The exploded view of the
assembly is Shown in Fig. 7.60. The index plate is attached to the angle plate using a
socket head screw. The indexing pin in screwed on the index plate.

Method of indexing The workpiece is located and clamped in position as discussed


in the previous problem. After the first hole is drilled, the clamping knob is loosened.
The workpiece is then rotated in such a way that the drilled hole is brought in align-
ment with the indexing pin. The indexing pin in the screwed forward so that the pin is
inserted in the drilled hole. The workpiece is thus locked in position for the next hole
to be drilled, and is then clamped. This process is repeated for all the holes.
Design of Drill Jigs 523

Fig. 7.59 Solid model of the Fig. 7.60 Exploded view or the designed drill jig
designed drill jig

Problem 4 For the workpiece shown in the previous problem, design an indexing
drill jig so as to drill two, four or eight equally spaced 10 mm holes of 15 mm depth at
a distance 30 mm form any end of the workpiece.

Solution: The workpiece is required to be located and clamped for drilling two, four
or eight equally spaced 10 mm holes of 15 mm depth at a distance 30 mm form any
end of the workpiece. This essentially requires designing an indexing drill jig. The
locating and clamping methods remains same as that discussed in the previous prob-
lem. However, in order to drill 2, 4 or 8 holes on the circumference of the workpiece,
it requires incorporation of an indexing mechanism. The indexing mechanism should
be capable of dividing the workpiece in 2, 4, or 8 equal divisions (i.e., 180°, 90°, 45°
divisions). For this purpose, a spring loaded indexing pin is used. The indexing pin is
inserted on the indexing holes provided on the stud, which locks the workpieces in the
desired position. The designed parts of the drill jig are presented in the Table 7.4 with
reference to the solid model and detailed drawing of individual part.

Table 7.4 Designed drill jig parts


SI. no. Part name/Part description Quantity Figure reference

1. Angle Plate 1 7.61 (a) and (b)


2. Bush Plate 1 7.62 (a) and (b)
3. Drill Bush 1 7.48 (a) and (b)
4. Socket Head Screw 3 7.49 (a) and (b)
5. Locating Stud 1 7.50 (a) and (b)
6. Clamping Knob 1 7.51 (a) and (b)
7. C Washer 1 7.52 (a) and (b)
8. Knob (stud rear end) 1 7.53 (a) and (b)
9. Indexing Pin 1 7.64 (a) and (b)
10. Compression Spring 1 7.65 (a) and (b)
11. Index Pin Knob 1 7.66 (a) and (b)
524 Tool Design

Fig. 7.61 Angle Plate (a) Solid model (b) Detailed drawing

Fig. 7.62 Index plate (a) Solid model (b) Detailed drawing
Design of Drill Jigs 525

Fig. 7.63 Stud (a) Solid model (b) Detailed drawing

Fig. 7.64 Indexing pin (a) Solid model (b) Detailed drawing

Fig. 7.65 Compression spring (a) Solid model (b) Detailed drawing
526 Tool Design

Fig. 7.66 Index pin knob (a) Solid model (b) Detailed drawing

The complete assembly of all the parts is shown in Fig. 7.67. The exploded view of
the assembly is shown in Fig. 7.68. The spring loaded indexing pin is housed in the
slot provided in the angle plate (refer to Fig. 7.61). The indexing pin is inserted in
the indexing slot provide in the stud (refer to Fig. 7.63) for locking the workpiece at a
particular position.

Method of Indexing The workpiece is lacted and clamped in position as discussed


in the previous problem. After the first hole is drilled, the stud rear end knob is loos-
ened. The indexing pin is then pulled out using the index pin knob. The workpiece and
the stud, both are then rotated in such a way that the next indexing slot is brought in
alignment with the indexing pin. The indexing pin in then released. The workpiece is
thus locked in position for the next hole to be drilled. The stud rear end knob is then
tightened. This process is repeated for all the holes.

Fig. 7.67
Design of Drill Jigs 527

Fig. 7.68 Exploded view of the designed indexing drill jig

Problem 5 Design a drill jig to drill four 10 mm through hole at PCD 30 mm on the
workpiece as shown in Fig. 7.69 (a) and (b).

Fig. 7.69 The finished workpiece (a) Solid model (b) Detailed drawing

Solution: The given problem requires designing a drill jig, for drilling four 10 mm
through hole at PCD 30 mm on the workpiece. The axial hole of 25 mm on the work-
piece can be conveniently used for location The workpiece can be located using a
stud locator and can be clamped using a C washer and knob. For this purpose a plate/
turnover jig can be used.
The designed parts of the drill jig are presented in the Table 7.5 with reference to the
solid model and detailed drawing of individual parts.
528 Tool Design

Table 7.5 Designed drill jig parts


SI. no. Part name/Part description Quantity Figure reference

1. Bush Plate 1 7.70 (a) and (b)


2. Jig Legs 4 7.71 (a) and (b)
3. Washer 5 7.72 (a) and (b)
4. Hex Nuts 5 7.42 (a) and (b)
5. Drill Bush 4 7.73 (a) and (b)
6. Locating Stud 1 7.74 (a) and (b)
7. Clamping Knob 1 7.51 (a) and (b)
8. C Washer 1 7.75 (a) and (b)

Fig. 7.70 Bush Plate (a) Solid model (b) Detailed drawing

Fig. 7.71 Jig Leg (a) Solid model (b) Detailed drawing
Design of Drill Jigs 529

Fig. 7.72 Washer (a) Solid model Fig. 7.73 Bush (a) Solid model
(b) Detailed drawing (b) Detailed drawing

Fig. 7.74 Locating stud (a) Solid model (b) Detailed drawing

Fig. 7.75 C Washer (a) Solid model (b) Detailed drawing

The complete assembly of all the parts is shown in Fig. 7.76 (a) and (b). The exploded
view of the assembly is shown in Fig. 7.77. The Jig can be turned over on its legs for
loading and unloading of the workpiece. The workpiece can be located using a stud
locator and the bush plate. After placing the workpiece on the stud, it is clamped using
530 Tool Design

a C washer and a knob. The stud is attached to the bush plate sing a hex nut at the
rear end. The jig legs are attached to the bush plate using hex nuts. The bushes are
fitted to the bush plate at desired locations, so the drilled hole will be located at the
correct position. The bush plate is designed for adequate swarf clearance between
the workpiece and the bush plate (refer to Fig. 7.70).

Fig. 7.76 Solid model of the designed drill jig (a) Showing the frilling side (b) Showing the
clamping side

Fig. 7.77 Exploded view of the designed frill jig

Summary
The process of hole making is an integral and difficult part of manufacturing. The
requirement of accurate size hole at an accurate location an in large number of parts
further enhances the complexity of the process. Jigs are suitably applied for this
purpose.
• A jig may be defined as a device which holds and locates a workpiece and
guides cutting tools.
• The holding of the work and guiding of the tool are such that they are located
in true positions relative to each other. Jigs are thus provided withy tool guid-
ing elements such as hardened steel drill bushes, which direct tool to correct
position on workpiece.
Design of Drill Jigs 531

• Jigs are employed for machining processes like drilling, reaming or tapping of
holes.
• Jigs are normally not bolted to the machine table and are provided with feet,
preferably four, opposite all surfaces containing guide bushings, so that it will
‘rock’ if not standing square on the table and so warn the operator.
• Jigs are provided with ample chip clearance for metal filings removal and clean-
ing. Clearances are also provided for overshoot of the drill.
• Jigs should be light weight to minimise operator fatigue, due to repeated
handling.

Questions
1. What is a drill jig?
2. What is the difference between a drill jig and a fixture?
3. Why is it poor practise to locate drill bushings in the leaf of a leaf-type drill jig?
4. Why is it considered poor practise to clamp the part with the hinged plate of a
leaf-type jig?
5. What is the advantage of tumble jigs?
6. Why is chip formation with a standard two-flute twist drill unlike that of other
metal cutting tools?
7. Why do standard twist drills require large axial thrust forces when drilling?
8. Referring to Question 7, how are the large axial thrust forces reduced?
9. What alterations to drill geometry will cause metal chips to break?
10. Why will increasing the drill feed sometimes cause chips to break on borderline
materials?
11. What is the difference between the major and burr formed during operations?
12. What problems are caused by major and minor burrs?
13. Why should small projections that extend outside the body of a drill jig be
avoided?
14. Why should complicated clamping devices be avoided?
15. Why should the use of loose parts on drill jigs be avoided?
16. Why should locating points on a drill jig be visible whenever possible?
17. What is the general rule for the distance between the bottom of a drill bushing
and the workpieces?
18. What are the three ways chips are removed from drill jigs?
19. Why is it often undesirable to use compressed air as a chip-removal
medium?
20. What is the purpose of providing area relief below the workpieces?
21. Why should a drill jig stand on four legs rather than three?
22. How should jig feet be placed in relation to the drill bushings?
23. Why should jig feet be ground after they are mounted on the jig base?
24. What is the hardness of standard drill bushings?
25. What is the general rule for the length of drill bushings?
26. What is the general rule for clearance between the drill and the bushing hole?
27. What is the ANSI classification of drill bushings, and how is each classification
used?
28. What methods are used when two or more holes are so close together that
it is impossible to have an individual standard bushing for each hole in the
workpieces?
29. What is the major difference between reaming with guide bushings and drilling
with guide bushings?
532 Tool Design

30. What is meant by built-up construction if a jig, and what is the advantage?
31. Why should dowel holes be drilled completely through both the jig component
and jig body?
32. Why is it necessary to stress-relieve jig bodies that have been constructed by
welding?
33. What is the advantage of a cast-constructed jig-body?

Design Problems
1. Design a drill jig to drill the 1.5 mm hole in the clevis pin shown in Fig. 7.78a.
Use the design procedure explained in Chap. 1.
2. Design a drill jig to drill the four 6 mm holes in the gland shown in Fig. 7.78b.
3. Design a drill jig to drill the 6 mm hole in the pin fork shown in Fig. 7.78c.
4. Design a drill jig to drill the 3 mm holes in the disk shown in Fig. 7.78d.
5. Design a tumble jig to drill all the holes in the workpieces shown in Fig. 7.78e.
6. Design a drill jig to drill the 20 mm holes in the clevis pin shown in Fig. 7.78f.
7. Design a drill jig to drill the five 6 mm holes in the bottom of the radio cabinet
shown in Fig. 7.79. The material is urea-formaldehyde plastic.
8. Design a drill jig to drill and ream the 16 mm holes in the shifter fork shown in
Fig. 7.80a.
9. Design a drill jig to drill and tap the M8 × 1.25 threaded hole in the artillery shell
shown in Fig. 7.80b.
10. Design a drill jig to drill and ream the holes in the workpiece shown in Fig.
7.81a.
11. Design a drill jig to drill the five holes in the crank arm shown in Fig. 7.81b.
Design of Drill Jigs 533

Fig. 7.78 Jig design problems


534 Tool Design

(b)

Fig. 7.79 Jig design problems: (a) Back view (b) Bottom view of radio cabinet
Design of Drill Jigs 535

M 8 × 12.5 M 52 × 8
16

75 R 55 52.5
48 60
R8 R 530

5 32
205
215
(b)

Fig. 7.80 Jig design problems: (a) Shifter fork (b) Artillery shell
536 Tool Design

R3
R 1.5
22 10

R3
6
R 16
22 Drill,
3 Holes

R6
16 2.5 Drill,
2 Holes
15°
82

R 16 45
32
33 R 16

32 16

R 16
95
(b)

Fig. 7.81 Jig design problems: (a) Governor fly ball (b) Crank arm

References
Books
• ASTME: “Handbook of Fixture Design,” F W Wilson (ed.), McGraw-Hill, New
York, 1962.
• “Tool Engineers Handbook,” 2b ed., F W Wilson (ed.), McGraw-Hill, New York,
1959.
• Colvin, F H and LL Haas: “Jigs and Fixtures,” 5th ed., McGraw-Hill, New York,
1948.
• Grant, H E: “Jigs and Fixtures,” McGraw-Hill, New York, 1967.
• NYSVPAA: “Jig and Fixtures Design,” Delmar Publishers, Albany, N Y, 1947.
• Sedlik, H: “Jigs and Fixtures for Limited Production,” SME. Dearborn, Mich.,
1971.
Catalogues
• Accurate Bushing Company, Garwood, NJ.
• Ace Drill Bushing Co., Los Angles.
• American Drill Bushing Company, Chicago.
• Carr Lane Mfg Co., St Louis, Miss.
• Ex-cell-o Corp., Detroit.
• Universal Engineering Co., Frankenmuth, Mich.
8 DESIGN OF FIXTURES

8.1 INTRODUCTION
A fixture is a device for holding a workpiece during machining operations. The name is
derived from the fact that a fixture is always fastened to a machine or bench in a fixed
position. It does not contain special arrangements for guiding the cutting tool, as drill
jigs do. In a set-up using a fixture, the responsibility for accuracy depends upon the
operator and the construction of machine tool.
Other types of tooling used for positioning parts relative to each other for fabricat-
ing purposes are also commonly referred to as fixtures. Assembly fixtures and weld
fixtures are examples of this type.
Many machining operations can be performed by clamping the workpiece to the
machine table without using a fixture, especially when a few parts are to be machined.
However, when the number of parts is large enough to justify its cost, a fixture is gen-
erally used for holding and locating the work.

8.2 FIXTURES AND ECONOMICS


A fixture for production machining must be considered like any other machine tool. It
should pay for itself from the savings derived from its use. The tool designer may be
required to determine the cost of a fixture and whether its construction will be profit-
able in the early stages of fixture design. The cost and return cannot be determined
exactly because of the many variables that may enter into the overall picture, but a
close estimate can be made from formulas developed for this purpose.
The economic aspects of the use of a fixture can be considered on the following
basis:1
1. The production per year necessary to pay out of savings for a fixture of given
estimated cost.
2. The maximum allowable cost of a fixture to break even for a given saving in
operating cost per piece and specific number of parts produced during a year.
3. The number of years it will take to pay out of savings for the proposed fixture
with a given saving in operating cost per piece and a given yearly production.
4. The net savings from the use of fixture of a given cost, for an estimated saving
in operating cost piece at a given output.
These various points are illustrated in the examples which follow. All the calculations
are based on the yearly rates; these rates should always be used in making calcula-
tions, even though production run may actually last less than one year.

1
From “A Treatise on Milling and Milling Machines,” 3d ed., Cincinnati Milacron, Inc., Cincinnati, Ohio, 1951,
by permission.
538 Tool Design

The number of pieces N which must be processed in a year


Example 8.1 to break even on a proposed fixture costing C dollars can be
obtained from the simplified formula
C (i + u + t + 1/a) + Si
N = __________________ ...(1)
s(1 + L)
where N= number of pieces produced per year
C= cost of fixture, dollars
S= difference between unamortised and scrap value of old fixture,
dollars
i= yearly interest on capital invested, expressed as a decimal
u= yearly percentage of cost for upkeep, expressed as a decimal
t= years percentage of cost for taxes, insurance, etc., expressed as a
decimal
a= years required for amortising the cost of the fixture out of savings
s= savings in direct operating cost per piece, dollar
L= percentage of burden, expressed as a decimal
Solution: Obsolescence due to change in product design is usually the predominant
factor in the life of a fixture, except where design changes occur less frequently. In
such a case, deterioration of fixture due to wear may predominate over obsolescence.
The number of years during which a fixture may be paid for out of savings will vary
with the stability of design of the parts processed in the shop. This period is often one
or more years; one year is the generally accepted figure, although deviations from this
figure may be round in the same shop.
If the selected values of i, u, t, and a are assumed as constant for a given period of
time, their combined value may be represented by a fixed charge factor f, so that
Eq. (1) can be further simplified, as follows:
Cf + Si
N = _______ ...(2)
s(1 + L)
Assume the following values: C = $400, i = 0.06, u = 0.04, t = 0.10, a = 2 years. Then
f = 0.06 + 0.04 + 0.10 + ½ = 0.70, s = $0.03, L = 0.50, and S = $100.
From Eq. (2), it follows that
(400) (0.7) + (100) (0.06)
N = _____________________
(0.03) (1 + 0.5)
= 286/0.045 = 2,860/0.45
= 6355 pieces per year
Thus, it can be seen that a fixture costing $400 can be paid for out of savings if
approximately 6355 pieces are produced per year.

For a given number of parts processed per year and known


Example 8.2 values of the fixed charges and operating cost, the maximum
allowable cost of a fixture to be paid for out of savings can be
calculated by solving Eq. (2) for C:
Ns(1 + L) – Si
C = ____________ ...(3)
f
Solution: Assuming the same values as in Example 8.1 but changing the number N of
parts to be processed to 5000 per year gives
Design of Fixtures 539

(5000) (0.03) (1.5) – (100) (0.06)


C = ___________________________
0.7
225 – 6
= _______
0.7
= $312.85
In order to break even in this case, the cost f the fixture should not exceed $312.85.
For very large production runs, Eq. (3) shows that it would be economical to make a
substantial investment in fixture, even through the saving in direct operating cost per
plece may be small.
The number of years a required to pay for a proposed fixture
Example 8.3 costing C dollars, when producing parts at a yearly rate N, can
be obtained from the formula
C
a = ______________________ ...(4)
Ns(1 + L) – Si – C(i + u + t)
Solution: If C = $600, N = 8000 pieces, and the other values are the same as in
Example 8.1, then
600
a = _________________________________________________
(8000) (0.03) (1.5) – (100) (0.06) – (600) (0.06 + 0.04 + 0.10)
600
= _________
360 – 126
600
= ____
234
ª 2.5 years
Find the net yearly savings on operations with a fixture of given
Example 8.4 cost and under given operating conditions. The net saving P of
a fixture is the difference between the gross operating savings
and the fixed charges:
P = Ns(1 + L) – Si – Cf ...(5)
Solution: If the number of parts N actually produced per year is 10,000, the fixture cost
C is $500, and other values are the same as in Example 8.1, the net yearly saving is
P = (10,000) (0.03) (1.5) – (100) (0.06) – (500) (0.7)
= 450 – 356
= $94 per year
This saving is approximately 20 per cent of the fixture cost.

8.3 TYPES OF FIXTURES


Examples of common types of fixtures are vise, milling, boring, broaching, lathe, and
grinding. Each will be described in the following sections:

8.3.1 Vise Fixtures


Standard machine vises adapted with special jaws provide an easy way of holding
parts for machining. They may be used with various types of machine tools and there-
fore may be classed as a type of fixture by themselves.
540 Tool Design

All machinists know that it is very easy to clamp a workpiece with parallel sides in
a vise but that worpieces with round or irregular contours are very difficult to clamp
properly. It is also well known that unless precautions are taken, the clamping action
of vise jaws is likely to mar finished surfaces. Special jaws are designed to hold work-
pieces with irregular contours properly and at the same time avoid damaging impor-
tant surfaces.
Figure 8.1 shows a series of special jaws. In a is shown a simple pair of jaws for hold-
ing a round workpiece, and in b a pair of jaws for holding a thin sheet of nonmagnetic
material. If the wrokpiece is steel, a magnetic fixture may be used. Note the stop
pin, needed to prevent bending or springs the wrokpiece by the clamping action. In
c, extended jaws for large workpieces are for securing alignment. The tongue and
groove in d assure accurate alignment.

Special jaw Workpiece Regular jaw 5° Workpiece

Vise Stop pin

(a) (b)
Extanded jaws

Bushing Guide pin Tongue Groove

(c) (d)
Direction of cut

(e) (f)

Fig. 8.1 Special vise jaws

When it is necessary to hold the workpiece firm against pressure in all directions, the
wedge-type jaws shown in e may be useful, and where the pressure exerted by the
Design of Fixtures 541

cutting tools is likely to cause one end of the workpiece to tilt upward, the link con-
struction shown in f should be used. This link construction has the further advantage
of permitting considerable variation in the dimensions of the workpiece, thus making
this construction suitable for handling rough castings and forgings.
All the jaws considered up to now have held the workpiece more or less in vertical
position. However, many workpieces require surfaces machined at a specified angle
other than 90°. Such a workpiece is shown in Fig. 8.2. One of the jaws is recessed
to take the workpiece at the proper angle. A locating stud is provided for the hole
previously drilled in the workpiece. A swing jaw assures the rigidity of the workpiece
without having to tighten the vise excessively.

Cutter outline

Swing jaw Swing jaw Workpiece


(soft stock) lift lever
Jaw Jaw Swing jaw Swing jaw
pivot stud stop pin

Workpiece Workpiece
locating stud stop pin

Fig. 8.2 Special vise jaws

Small castings or forgings with irregular contours can be held in standard machine
vises by using epoxy resin to form a nest within the vise jaws to fit the irregular con-
tour. This type of tooling is explained in detail in Chapter 11.

8.3.2 Milling Fixtures


A mill fixture holds the part in the correct
relation to the milling cutter as the table
movement carries the part through the
cutters. It consists of five main parts,
the base, clamps, rest blocks or nest,
locating points, and gauging surfaces
(Fig. 8.3).
The base of a mill fixture consists of a
baseplate which has a flat and accu-
rate undersurface and forms the main
body on which various components are
mounted. This surface mates with the
surface of the mill table and forms the
reference plane with respect to the mill Fig. 8.3 Main parts of a mill fixture
feed movements. It may be constructed (Cincinnati Milacron)
542 Tool Design

of steel plate or cast iron, depending upon the size and complexity of the part. Steel
plate may be in the form of AISI 1020 hot-rolled or cold-rolled plate. Cold-rolled plate
is closer to dimensions and has a better finish. It is used when the base requires little
machining and when the fixture is held together mechanically. When the base is fab-
ricated by welding, cold-rolled steel has a tendency to warp more than hot-rolled plate
because of the stresses set up during the rolling operation. Therefore, it is better to
construct welded-fixture bases of hot-rolled steel plate, since welding and the follow-
ing stress-relieving treatment will scale the plate anyway. It is also better to construct
the base of hot-rolled steel when considerable machining is done, as hot-rolled steel
plate is lower in cost.
The base is provided with slots for the pur-
pose of clamping the fixture to the mill table.
It also has a keyway running lengthwise in
the base for two keys used to align the fix-
ture on the milling-machine table. The keys Hold-
are pressed into the keyway at both ends down
of the fixture and held there by socket-head slots
cap screws as shown in Fig. 8.4.
Key
Before staring the design of a mill fixture, it is
quite important to know dimensional data of
the milling machine on which the fixture will
be used. These data include the dimensions
of T slots, the centre-to-centre distance of Fig. 8.4 Underside of mill-fixture base
the T slots, the dimensions of the milling-
machine table, and the length of table travel in all three feed movements.
Referring to the above, it will be necessary to know the dimensions of the T slots and
the centre-to-centre distance of the T slots when designing the base shown in Fig. 8.4.
It is also important that the tool designer provide enough clearance space around the
hold-down slots for a nut, washer, and wrench. If the fixture is to be used on more than
one machine, it may be well to hold the fixture with strap clamps as shown in Fig. 8.5,
because T-solt centre-to-centre distances vary from machine to machine.

Strap clamp

Fixture base

Mill table

Feed movement

Fig. 8.5 Use of strap clamps to hold fixture to machine table

The principles of clamps and clamping on a mill fixture are the same as for other types
of tooling, except that clamps on mill fixtures must be extremely rigid. Cutting forces
Design of Fixtures 543

exerted by a milling cutter may change as the cutter enters or leaves the work and
throw an extra load on clamps (see Fig. 8.6). Clamps must not loosen by vibration
caused by interrupted cutting of the mill cutter. Interrupted cutting occurs especially at
the beginning and end of the cut.
Clamps should be located opposite bearing surfaces and locating points and should
be so designed that they can be easily manipulated by the mill operator. They must
be in a position to clear the milling cutters easily.

Force

Work Work
Force
(a) (b)

Fig. 8.6 Effect of cutting forces on workpece at beginning of cut (a) Conventional milling:
work is lifted up at beginning of cut (b) Climb milling: work is forced down at one
end but lifted up at the other

The rest blocks or bearing surfaces are located within the nest and provide support for
the workpiece. These surface vary in design according to the shape and size of the
workpiece but are usually in the form of pins, pads, or plates that are accurately placed
in the base of the fixture (see Fig. 7.18 for standard types). These bearing areas are
raised above the surface of the base to permit chips to fall away and allow easy clean-
ing. Pins, pads, and plates should be heat-treated when subjected to heavy use.
For very accurate work, the accuracy of contact between the bearing surface and the
workpiece resting upon it may be checked by recessed reference surfaces as shown
in Fig. 8.7. Several reference points are placed around the part. A feeler gauge of pre-
determined thickness is inserted between the surfaces, and when the bearing contact
is obtained the feeler gauge will fit snugly.

Fig. 8.7 Methods of checking bearing surface between workpiece and rest pads
544 Tool Design

As previously described, the purpose of a milling fixture is to hold the workpiece in


the correct relation to the cutter. The fixture is attached to the milling-machine table
and held in alignment by two keys attached to the fixture base. These keys fit very
closely into the t slot of the mill table. The fixture is held firmly to the table by T bolts
or hold-down clamps.
It is then necessary to adjust the table by use of feed movements until the correct
position is attained. This may be done by trial-and-error cuts in the workpiece, but it
is usually accomplished with gauges or properly located setting surfaces, which may
be set permanently in the fixture or may be removable if they are in the path of the
cutter.
Removable setting surfaces should be avoided whenever possible as they are more
subject to error and expensive to make.
A common practise is to mount feeler surfaces permanently in the fixture base or
body at a predetermined distance (usually 3.2 mm) below the proper cutter setting, as
shown in Figs 8.8 and 8.17. A feeler gauge is then used to determine when the cutter
is in the correct position. This avoids damage or wear on these surfaces because they
are set by the thickness of the feeler gauge from the cutter.

Fig. 8.8 Use of setting gauges to locate mill fixture in correct relationship to cutter

Feeler surfaces should be marked with the correct size of the feeler-gauge thickness
to aid the operator in set-up. Sometimes a groove is milled around the feeler surface
and filled with bright paint to note its location. This is to attract the attention of the
operator in order to prevent accidental surface damage by the cutter.
These are differences of opinion whether feeler surface should be hardened or left
soft. One school advocates that they be soft. The belief is that if the cutter accidently
Design of Fixtures 545

hits the feeler surface, it is more economical to replace the surface than to reshapen
or replace the cutter, as would be the case if the cutter hits a hardened surface. The
other school believes that there is a possibility of the operator’s milling completely
through the feeler surface without being aware of it. This, of course, would destroy the
reference surface, and it would not be discovered until part dimensions are checked.
If the surface is hardened, the operator would certainly know when the cutter strikes
the reference surface.
A word of caution with regard to setting gauges. They work well in theory, but there are
too many variables that enter into a milling operation to allow an operator to depend
upon them entirely. For instance, cutter runout or one high cutter tooth may cause
an extra amount of material to be removed. Accuracy in setting with feeler gauges
depends upon the sense of feel, which may vary between operators. The first parts
should be checked immediately after the set-up and allowances made for variables
that enter into a milling operation.

Classification and types of mill fixtures Mill fixtures are classified in a variety of
ways, and there seems to be no standard method of classification. This is because
each fixture is different and must be designed to meet certain requirements. The tool
designer will use his past experience and knowledge when determining the general
type of fixture to use. He may combine features of two or even three general types of
fixtures to make the fixture meet the necessary requirements. A considerable amount
of forethought should go into the design of fixture before deciding on a certain type.
Before discussing the various classifications and types of fixtures it would be well to
consider the various methods of producing milling. The general methods are single-
piece, string, reciprocal, progressive, index, and rotary milling (see Fig. 8.9).

Fig. 8.9 Various methods of production milling: (a) Single-piece (b) String (c) Reciprocal
(d) Progressive (e) Index (f) Rotary milling
546 Tool Design

Single-piece milling One workpiece is held in a vise or fixture and fed through the
milling cutter. The mill table is returned to its starting point before inserting another
workpiece.

String milling A series of identical workpiece is mounted in line parallel to the table
feed movement and fed in consecutive order into the mill cutter. The workpiecces
should be as close together as possible in order to keep dead time between the
parts to a minimum. Sometimes two rows of workpieces are located side by side or
abreast.

Reciprocal milling Fixtures are mounted on each end of the mill table with the mill
cutter in the middle. The operator unloads and loads one workpiece while the other is
being machined.

Progressive milling Two or more milling operations are performed simultaneously


on separate but identical workpieces. One may be completed with each pass of the
mill table, but it is necessary to move a partially finished part to the next station in the
fixture and insert a new part in its place.

Index milling Duplicate fixtures are mounted on a table which is pivoted in such a
manner that it can be rotated and indexed into position. This has the advantage that
the operator can stand or sit in one position and do the reloading at one station. The
disadvantage of index milling is that there may not always be sufficient clearance to
swing the table.

Rotary milling A series of fixtures is mounted on a large rotary table which in turn
rotates the workpiece consecutively under the cutter. The table must be large enough
to give the operator ample time to remove and insert a new workpiece without stop-
ping the table. For this reason, the fixtures must be easily cleaned, contain quick-
acting clamps, and have a positive locating system.
Mill fixtures used in the foregoing milling operations are classified in a variety of
ways: (i) According to the types of milling operation performed on the work, such as
face-milling fixtures, slab-milling fixtures, slotting fixtures, straddle-milling fixture, etc.;
(ii) According to the way the workpiece is clamped, such as hand-clamping fixtures,
power-clamping fixtures, toggle fixtures, or automatic fixtures, the clamping and load-
ing of which is arranged to function in conjunction with the cycle of the machine;
(iii) According to the way the workpiece is located, such as centre fixtures, V-block
fixtures, and pin or stud fixtures.
Mill fixtures can also be classified by the method of presenting the workpiece to the
cutter, as in cradle fixtures. Here the workpiece is rocked or rotated within a given
angle during milling. Rotary fixtures, where the workpiece is rotated under the cut-
ter, and drum fixtures, where the workpiece is mounted on the periphery of a rotat-
ing drum, fall under this classification. Also included are indexing fixtures, where the
workpiece is indexed into the next position during the machining cycle of the mill, and
rise-and-fall fixtures, which allow raising and lowering of the workpiece is conjunction
with the mill feed.
Design of Fixtures 547

Other general types of mill fixtures are universal fixtures, designed to hold a family
of parts similar in shape but different in size, and progressive fixtures, where the
workpieces can be located in different positions and progressively moved from one
fixture station to the next until all operations are complete. Temporary mill fixtures
are built of standard clamps, T bolts, heel blocks, tacks, and stops, and the table of
the mill is used as the base of the fixture. Such a set-up is usually used in toolroom
operations or small-lot production (see Figs 8.10 to 8.17 for example of various types
of mill fixtures.)

Fig. 8.10 Hydraulioc indexing fixture utilising hydraulic clamping methods (Cincinnati
Milacron, Inc.)

Fig. 8.11 Temporary indexing fixture utilising a circular milling table (Cincinnati Milacron,
Inc.)
548 Tool Design

Fig. 8.12 Face-milling fixture showing manual clamping methods (Cincinnati Milacron,
Inc.)

Fig. 8.13 Cradle fixture with hydraulic Fig. 8.14 Temporary fixture using
clamping and indexing methods various clamps, step blocks,
(Cincinnati Milacron, Inc.) and jacks (Cincinnati
Milacron, Inc.)
Design of Fixtures 549

Fig. 8.15 Rotary fixture utilising Fig. 8.16 Seven-station rotary fixture
hydraulic clamping methods (Cincinnati Milacron, Inc.)
(Cincinnati Milacron, Inc.)

Fig. 8.17 Vise jig and cutter-setting gauge (Cincinnati Milacron, Inc.)

Obviously these fixture classifications may overlap. This should not be of great con-
cern to the tool designer, but he should be familiar with the various general types and
how they overlap.
The discussion of milling fixtures would not be complete without referring to the use
of magnetic and vacuum chucking methods. Magnetic clamping of workpieces has
long been a common practise for surface-grinding operations, and its advantages as
a work-holding method have long been known. Recent developments in the design
of magnetic chucks have enabled many plants to use them for production milling
operations.
550 Tool Design

For example, the Sundstrand Machine Tool Company has developed a low-voltage
chuck with a great deal more holding power because of the increased efficiency of
the magnetic path within the chuck itself. This is accomplished through a reduction
of internal insulation. The steel members of the magnet are closer to the energising
coil forces, which results in more available holding power with lower input voltage.
The chucks are operated with a power unit which converts 110/220 or 440-volt ac
to 6-, 12-, or 24-volt dc. The low-voltage feature ensures safety to the operator and
provides heat-free holding, which minimises distortion of the workpiece with resulting
improvement in final accuracy.
Figures 8.18 and 8.19 show examples of milling operations using magnetic chucks as
holding devices. In Fig. 8.18, the cut is 3 mm deep, with a 152 mm diameter carbide-
tipped face-milling cutter being used at 272 rpm and feed at 0.4 m per min. In this
case, the chuck is magnetised and demagnetised automatically with the movement of
the machine table. Note that the cutting forces are directed toward a fixed stop.

Fig. 8.18 Use of magnetic chuck for milling Fig. 8.19 Use of two-way magnetic
rough irregular casting (Sundstrand chuck for slot milling (Sunds-
Machine Tool) trand Machine Tool)

In Fig. 8.19, two-way magnetic clamping makes it possible to mill 2.4 mm slots 12 mm
deep in 14 steel nuts per cycle, 350 pieces per hour were produced with this set-up.
Vacuum chucks are being used for holding nonferrous and nonmagnetic parts for
milling operations where holding without distortion is vital. A vacuum chuck is a suc-
tion holding device relying on a vacuum pump to reduce the absolute pressure in a
chamber to a pressure below atmospheric. The part being chucked is positioned over
this chamber and is acted upon by atmospheric pressure, which exerts a downward
force on all surfaces exposed to vacuum. The requirements of materials to be held by
vacuum are that sufficient area be provided for atmospheric pressure contact (mini-
mum 13 cm2), that the surface contacting the vacuum top plate be reasonably flat,
and that the materials be nonporous.
The vacuum chuck shown in Fig. 8.20 is based on a unique vacuum valve system.
A pattern of tapped ports, fitted out with individual valve screws, permits controlling
the vacuum area merely by opening each valve screw one turn directly under the part
Design of Fixtures 551

to be chucked. All valves not in use are sealed by individual O rings. Thus, control-
ling the valve screws makes possible the chucking of regular or irregularly shaped
parts. Figure 8.21 shows the vacuum chuck in use on a numerically controlled milling
machine. Note that side stops position the part over the vacuum area.

Sealed Open

Atmospheric pressure up to Top plate


“O” ring seals in vacuum
12 lbs. PSI stop
Atmospheric ports seals out coolant
Chucked
part Side rail
Side rail

To vacuum
pump
Vacuum gauge Atmosphere enters
ports to release part
Push valve to Cast iron Mounting
release part Adjust knurled nut to chuck body tongue
vary holding power

Fig. 8.20 Universal vacuum chuck (Dunham Tool Company, Inc.)

Fig. 8.21 Use of vacuum chuck for holding part on numerically controlled milling machine
(Dunham Tool Company, Inc.)
552 Tool Design

Magnetic and vacuum chucks can be more economical because in many cases the
need for expensive mechanical, hydraulic, or air clamping devices is eliminated. This
may result in the reduction of the machine table needed to hold a given number of
parts or permit more parts per set-up. Moreover, smaller-diameter cutters may be
used since no clearance is required, as with mechanical clamps.

8.5 BORING FIXTURES


Boring fixtures may have the characteristics of a drill jig or a mill fixture, depending
upon the types of boring operation. The workpiece always has an existing hold which
is enlarged by the boring operation. It may be the final operation, or it may be prelimi-
nary to grinding and other sizing operations.
Boring is usually accomplished with a single-point tool, and the size of the hold
depends upon the adjustment of the tool within the boring bar. Boring will correct hole
location and out-of-roundness, as the tool removes more material from one side of the
hole than the other and does not follow the axis of the original hole. It can be accom-
plished by rotating the bar or by rotating the workpiece around the bar.
Machines used for operations are drill presses with power feed, milling machines,
lathes jig-boring machines, vertical or horizontal boring mills, and special-purpose
boring machines operated manually or automatically.
The principle of boring is basically the same on all these machines. The boring bar is
fixed and the workpiece moves into the bar, or the workpieces are fixed and the bar
moves into the wokpiece (see Fig. 8.22).
The general principles of fixture design apply in the design of boring fixtures. Boring
fixtures which are used on machines that have adjustment on the machine table or
boring bar may require additional indicating surfaces or holes for initial alignment.
Clearance should be provided for bar travel and boring-tool adjustment. The impor-
tance of clamping directly over supporting surfaces should be emphasised since pres-
sure exerted on an unsupported surface will induce a twist or bow in the work. The
result will be nonparallel holes when the clamping forces are released as shown in
Fig. 8.23.
Boring fixtures are divided into two general classes. In one, the fixture guides the bor-
ing bar, as in drill jigs; in the other, the fixtures holds the work in the proper relation to
the bar, as in mill fixtures. The particular class of boring fixture to be used depends
upon the type of boring bar, the type of boring machine, or both.
Boring bars are generally classed as a stub bar, a single-piloted bar, and the dou-
ble-or multiple-piloted bar. Stub bars are commonly used for boring short holes, as
in jig-boring operations. They are also used to bore blind holes that do not permit the
extension of a bar through the work. Boring fixtures for stub bars serve to hold the
work in the correct relation to the bar. Multiple holes are bored by moving either the
work or boring bar (depending upon the type of machine) through x and y coordinates
from a zero reference point.
A single-piloted bar is longer than a stub bar and is guided on its leading end to pre-
vent springing under the cutter thrust as shown in Fig. 8.24. An end support is pro-
vided on some types of boring machines and may be readily brought into alignment
by adjustment of the bearing block. In this case, the purpose of a fixture again is to
Design of Fixtures 553

Fig. 8.22 Basic principles of (a) Horizontal and (b) Vertical boring operations

Fig. 8.23 Nonparallel holes as the result of clamping over unsupported surfaces.
554 Tool Design

hold the workpiece in the correct relation to the boring bar. However, it may be neces-
sary to make the support an integral part of the fixture, as shown in Fig. 8.25.

End Boring bar Boring


support tool

Work Boring machine


headstock

Machine
table

Fig. 8.24 Single-piloted boring bar

Machine spindle
Boring bar
Master hole
for indicating Workpiece
fixture
Boring tool Fixture

End support

Machine
table

Fig. 8.25 Line boring operation using end support mounted in fixture. Set-up is on drill press
with power feed

Where diameters of boring bars permit, they can be guided by standard drill bush-
ings fitted into the fixture. Guide bushings may also be made of Meehanite, hardened
steel, bronze, or bearing alloy, to suit the hardness of the bar and the length of the
production run. Anti-friction bearings may be used when production is high enough to
allow for this more expensive construction or when high bar speeds are required. The
wearing surfaces of the bar may be chrome-plated or hardened to provide additional
wear resistance.
The double-piloted bar is guided ahead and behind the cutter, as shown in Fig. 8.26.
The rear pilot is larger than the sweep of the cutter so that bar can be withdrawn
through the bushing. Since the bar is guided at both ends, a floating adaptor connec-
tion is provided between the bar and the machine spindle and is necessary to com-
pensate for small differences in alignment between the axis of the fixture bearings
and the machine spindle.
Design of Fixtures 555

Workpiece
Boring bar Slip bushing

Fixture
Floating
driver

Fig. 8.26 Double-piloted boring bar

Workpieces may be of a design that does not permit supporting the boring bar as
shown in Fig. 8.27. It may be possible to use the workpiece as support for the boring
bar. In the application shown in Fig. 8.27, the finished bore behind the cutter has been
bushed to the bar size. This type of design should be avoided whenever possible, as
it is time consuming to use and the alignment accuracy depends upon the previously
finished hole.

Cutting tool Workpiece


Spindle

Bushings inserted
in previously
bored holes

Machine table

Fig. 8.27 Boring-bar support by use of workpieces

The boring fixtures described so


far are generally used for small-lot
production or for relatively large
workpieces. Precision boring as a
long-run production operation is
usually done on a precision single-
point boring machine designed for
this purpose. Such a machine may
be vertical or horizontal and have
single or multiple spindles. There
are a variety of these machines on
the market in the small and medium
work ranges (see Figs 8.28 and
8.29). Fig. 8.28 Production single-point boring machine
used to bore and face lathe headstocks
(Heald Machine Division, Cincinnati
Milacron, Inc.)
556 Tool Design

Fig. 8.29 Production single-point boring machine used to precision-bore diesel connecting-
rod assemblies (Heald Machine Division, Cincinnati Milacron, Inc.)

This type of machine has little or no provisions for alignment of the part with the
spindle. The fixture is designed to mount on the machine table and present the work-
piece to the spindle or boring bar in the proper location for machining. Such a fixture
is usually referred to as a stationary fixture. The body of the fixture is usually a heavy
angle plate or platen or some variation of one or the other. Clamps for this type of
production fixture should be designed for quick and convenient loading. Provisions
for minor adjustment may be built into the fixture by means of jackscrews as shown
in Fig. 8.30.

Fig. 8.30 Method of providing minor adjustment in stationary boring fixture

If more than one stationary fixture is to be used on an individual machine, provisions


should be made for the removal and replacement of the fixture. The sides of the
Design of Fixtures 557

fixture bases should be machined parallel and/or at right angles to the direction of the
machine feed movements to allow the sides to fit against aligning rails provided for
this purpose on the machine table. Another method is to provide keys in the base of
the fixture which fit accurately in T slots of the boring-machine table, as is done on
the base of mill fixtures.
Other types of fixtures used on precision single-point boring machines are universal
fixtures (Fig. 8.31), indexing fixtures (Fig. 8.32), and automatic-loading fixtures (Fig.
8.33).

Fig. 8.31 Universal boring fixture (Heald Machine division, Cincinnati Milacron, Inc.)

Fig. 8.32 Indexing fixture used to finish-bore compressor disks (Heald Machine Division,
Cincinnati Milacron, Inc.)
558 Tool Design

Fig. 8.33 Automatic boring fixture on production boring machines (Heald Machine division,
Cincinnati Milacron, Inc.)

Universal fixtures may be obtained commercially from the manufacturers of boring


machines. Such a fixture consists of a base provided with a slide for cross movement
and adjustment and a vertical slide for vertical movement and adjustment. The slides
contain T slots or a series of tapped holes for mounting workpieces or supplementary
fixtures.
Indexing fixtures consist of a base or slide with a rotary table or rotating indexing
plate mounted on it. The table or plate may be provided with degree graduations or a
method of positive and accurate indexing. Such a fixture allows the operator to load
and unload the machine during operation.
Automatic loading fixtures are sophisticated examples of tool design. They may be
semi-automatic or fully automatic. Such a design is suitable only for long production
runs of a particular wokpieces. The tool designer will usually work with the manufac-
turer of the boring machine in the design of this type of tooling.

8.6 BROACHING FIXTURES


Broaching operations are generally classed as internal and external. General inter-
nal broaching, such as sizing and finishing of holes, does not require special fix-
tures as a rule. Fixtures for this purpose are often simple plate, plug, or horn fixtures
which established the correct position of the workpieces in relation to the broach and
machine faceplate. Figure 8.34 shows a simple fixture used for broaching a keyway,
the details of which should be studied.
Sometimes, it is necessary to cut a keyway deeper than can be achieved in one
pass. This is accomplished by inserting shims of the desired thickness (see Fig. 8.35).
Progressively thicker shims are inserted at the end of each pass until the required
depth has been reached. Shims also minimise wear on the broach and allow for
adjustment after the broach is sharpened. Both the shim and work horn should be
made of tool steel and heat-treated for maximum wear resistance.
Design of Fixtures 559

Machine faceplate
Plate adapter

Broach

Work horn

Workpiece

Fig. 8.34 Keyway-broaching fixture

Broach

Workpiece
Machine faceplate
Workpiece Work horn

Broach
Plate adapter

Shim Machine
faceplate
Work horn
Plate adapter

Fig. 8.35 Use of progressively thicker Fig. 8.36 Broaching keyway parallel
shims to complete keyway to side of taper bore
in more than one pass

Special work horns may be designed to broach keyways at an angle. Figure 8.36
shows an example of broaching a keyway parallel to the side of a tapered hole.
External broaching will usually require a special fixture for each job. Such a fixture
must withstand tremendous cutting pressure and therefore must be constructed rig-
idly enough to hold the part solidly during the broaching pass. The workpiece should
be supported by fixed stops and positive clamps. Cutting forces should be directed
toward fixed stops. It is important that the broached surface remain parallel to the
broach axis during the broaching operation. Any slipping or tilting of the part during
broaching could cause it to jam against the broach. Figure 8.37 shows an external
broaching fixture.
560 Tool Design

Shuttle broaching fixtures consist of a workpiece holder which is automatically with-


drawn from the cutting edges of two broaches for loading and then is returned for the
cutting operation. By this means the machine is made safer and easier to load and
unload. It also permits loading and unloading one fixture while the other is working
(see Figs 8.38 and 8.39).

Broach

Workpiece

Locating block

Fig. 8.37 External broaching fixture

Fig. 8.38 Single-ram vertical-surface Fig. 8.39 View of fixture on single-ram


broaching machine with vertical-surface broaching
shuttle table (LaPointe machines (LaPointe Machine
Machine Tool Company) Tool Company)
Design of Fixtures 561

High-production broaching may warrant the design and construction of index fixtures
on which the operator need only load the work. The unloading is automatic.

8.7 LATHE FIXTURES


A large majority of lathe operations can be accomplished by using standard chucks
and holding methods. However, many parts such as castings and forgings cannot
readily be mounted by any of the standard methods. It is therefore necessary to man-
ufacture special work-holding fixtures for machining these parts.
The basic principles of fixtures design apply to lathe fixtures. However, there are addi-
tional considerations that apply specifically to lathe fixtures, since the fixture and the
workpiece revolve. They are as follows:
1. Clamps and other holding devices should be designed in such a way that they
will not be loosened by centrifugal force.
2. The workpiece should be gripped on its largest diameter, and whenever possi-
ble the gripping diameter should be large than the diameter being machined.
3. The fixture should be well balanced. Irregular-shaped workpieces may have to
be counterweighted to prevent vibration. It is not necessary to achieve perfect
balance unless the workpiece is rotating at high speeds.
4. The fixture should be as light weight as possible it is rotating.
5. Projections and sharp corners should be avoided as the operator may not be
able to see them while the tool is rotating.
6. Thin sections of the workpiece may require support to resist the pressure of the
lathe tool.
Lathe holding methods are usually some adaptation of chucks, faceplates, and man-
drels or a combination of the three. Each of the foregoing methods has a standard
counterpart, such as three-jaw universal chucks, spring collets, rubber flex collets,
standard mandrels, etc. Such standard work-holding methods are universal, i.e., they
will hold a variety of work sizes as long as they are of the same shape or geometry.
Holding this general type of work is usually no problem and will not be discussed in
length. This section will be devoted to hard-to-hold workpieces and adaptations of
standard lathe holding devices.

Chucks Widely used lathe holding devices are universal self-centreing scroll chucks
with two, three, four, and six jaws and independent four-jaw chucks used for odd-
shaped workpieces. There are several variations, such as combination chucks with
self-centreing and independent jaw actions and jaws powered by air or hydraulic
cylinders on the rear of the spindle. The jaws, however, are all basically the same
design and are made in such a way that the outside section can be removed and
interchanged with special jaws, as demonstrated in Fig. 8.40. Such jaws are known
as top jaws.
Special top jaws may be designed for better fit to the gripping diameter of the work-
piece. Standard jaws usually will not be in full contact with the wokrpiece (see Fig.
8.41). Soft top-jaw blanks available from chuck companies can be machined to the
desired shape directly on the chuck. The jaw height should be kept short in order to
grip parts as close to the chuck face as possible. The finished jaw may be hardened
562 Tool Design

or left soft as desired. Hardened jaws with a medium 60° serration as shown in Fig.
8.42, are generally used for gripping rough work. However, on soft material, a medium
serration may penetrate too deep, and the surface will not clean up when machined. In
such cases, use a fine 60° serration, also shown in Fig. 8.42. Soft jaws or hard smooth
jaws are commonly used on previously finished or smooth diameters.

Soft top-jaw blank

Standard reversible
top jaw

Chuck
body

Special top jaw after


machining and heat
treating

Fig. 8.40 Standard and special top jaws

Fig. 8.41 Standard jaws and Fig. 8.42 Jaw serration dimensions: (a) Medium,
improper contact (b) Fine (Warner & Swasey Company)
with wokrpieces
Design of Fixtures 563

In addition to serrations, the gripping surfaces may be reinforced by the insertion of


hardened cone-point screws or pins, as shown in Fig. 8.43. The cone points sink into
the work ahead of the serrations and give a better grip on hard-to-hold surfaces, such
as draft angles on castings.

Fig. 8.43 The use of cone-point screws for extra grip (A) Heel thickness kept to a minimum
(B) Jaw height as short as possible; parts gripped close to chuck (C) On small-
diameter parts jaw gripping width must be watched to prevent pinching workpiece
between two jaws (D) This distance must always be kept less than one-half the
diameter of the workpieces for proper centreing action of jaws on work (Warner
& Swasey Company)

It is quite important when designing special jaws to be certain that the jaw grip-
ping surface is not too wide. Gripping surfaces that are wider than necessary have
a tendency to contact the workpiece at directly opposing points, which results in
improper centreing of the workpieces. Figure 8.44 shows how this situation can be
remedied by chamfering the corners of the jaws. The amount of chamfer depends
upon the diameter of the workpieces. A rule of thumb for the amount of chamfer is
given in Fig. 8.44.

Fig. 8.44 Chuck jaws with (a) Incorrect and (b) Correct gripping areas
(Warner & Swasey Compnay)
564 Tool Design

Using a set of rocking jaws on three-jaw chucks helps ensure a secure grip on the part
regardless of the surface roughness in the gripping area. Two of the three jaws are
solid with the third rocking. The solid jaws align the part, while the rocking jaw adjusts
to the uneven surface. The centre of the jaws should be undercut so that they grip
only on the ends (see Fig. 8.45).

Fig. 8.45 The use of horizontal rocking jaws (Warner & Swasey Company)

Figure 8.46 illustrates a solid set of jaws which grip long castings and forgings in a
manner similar to rocking jaws but cost less than rockers to make. They are suitable
for use on two-and three-jaw chucks. Two of the jaws chucks should have gripping
areas on each end. The gripping areas should not overlap.

Fig. 8.46 Solid jaws machined to grip like rocking jaws (Warner & Swasey Company)

Figure 8.47 shows a set of jaws which rock in a vertical plane. They grip on six equally
spaced areas, which distributes the gripping pressure more evenly. Such jaws are
especially suited to chucking thin-walled castings and forgings.
Design of Fixtures 565

Fig. 8.47 Vertical rocking jaws (Warner & Swasey Company)

Wraparound jaws are particularly useful on fragile parts that require gentle gripping.
They distribute the gripping pressure over more area, thereby minimising distortion
problems (see Fig. 8.48). The greater gripping area also provides more friction for
better drive. In cases where gripping pressure must be extremely light, a drive pin
may be provided in one jaw to pick up a cored opening or a protrusion on the part.
With a drive pin the jaws have only to centralise the part, as the pin provided the driv-
ing force.
The bore diameter of the wraparound jaws should be held to within 10.001 in. of the
turned diameter of the workpiece. Ths ensures an even distributior of the gripping
pressure over, the entire chuking area. If the bore diameter of the jaws is too big, a
three-point grip results; if too small, the jaws will dig in on the six corners.

Fig. 8.48 Wraparound jaws for gripping fragile or thin-walled parts


(Warner & swasey Company)

Faceplate fixtures A lathe-faceplate fixture is fastened to as standard lathe face-


plate and is usually built up from standard clamping and locating devices. It may be
similar in appearance to a mill fixture, but it is quite different in that it is rotating around
the lathe axis. The base o the fixture may be a flat plate bolted to the faceplate as
shown in Fig. 8.49, or it may be an angle plate, as shown in Fig. 8.50. It is usually
located on the faceplate by means of two dowel pins and secured by cap screws
566 Tool Design

inserted through the faceplate into tapped holes in the fixture or by T bolts inserted
into T slots in the faceplate. Sometimes, faceplate fixtures with circular bases have
indicating grooves or accurately machined holes for the purpose of aligning the fixture
by a dial indicator.

Fig. 8.49 Simple faceplate fixture

Fig. 8.50 Faceplate fixture for in-line boring

High-speed turning operations require that the fixture and workpiece be dynamically
balanced as accurately as possible. Low-and medium-speed turning operations do
not require such exact balancing methods, and reasonably accurate static balancing
is satisfactory.
A sliding counterpoise is sometimes used as a balancing device. Figure 8.51 shows a
type that is commonly used. It is particularly advantageous when several pieces of a
similar kind but of different weights are machined.
Figures 8.52 and 8.53 give two examples of faceplate fixtures in use. Figure 8.52
shows an angle types with a compressor cylinder block that requires boring the crank-
shaft bearings and facing the ends. Figure 8.53 shows a swinging-index types. The
operations to be performed on the workpiece are facing, drilling, boring, and tapping.
The swinging plate is swung to one end position and these operations performed. The
plate is then swung to the other end position, clamped, and the operation repeated.
Design of Fixtures 567

Fig. 8.51 Faceplate fixture showing use of adjustable counterpoise

Fig. 8.52 Angle-type fixture (Warner Fig. 8.53 Swinging-index-type fixture


& swasey Company) (Warner & swasey Company)

Magnetic and vacuum chucks The rotary magnetic and vacuum chucks shown in
Figs 8.54 and 8.55 are not lathe fixtures in the same sense as previous examples but
are unusual enough to warrant mention here. They are commercial items that can be
purchased from various companies and may save many hours of the tool designer’s
and toolmaker’s time.
Magnetic and vacuum chucks have the advantage of being able to hold thin and thin-
wall parts during turning and boring operations without distortion. It should be empha-
sised that they are suitable only for light turning operations, especially if there is a
568 Tool Design

small amount of bearing surface against the chuck face. This is of no real significance,
as parts held on these chucks are usually too fragile or thin to withstand heavy cuts.

Fig. 8.54 Vacuum-contoured chuck used to produce missile nose cone (Dunham Tool
Company, Inc.)

Mandrels Mandrels are shafts especially


made to hold work to be machined concen-
trically around a previously bored or drilled
hole. There are two general types in com-
mon use, plain and expanding.
Plain mandrels have taper of about 1:200.
Workpieces are pressed on the plain mandrel
by an arbor press. There must be a mandrel
for each hole size, and the tolerance must
be quite close on hole size to prevent the
mandrel from completely passing through Fig. 8.55 Magnetic lathe chuck (Brown
the workpiece. Mandrels are generally not & sharpe Mfg. Co.)
suited for production purposes, as they must
be removed from the machine to be pressed into the workpiece and there is consider-
able wear on the mandrel when it is pressed in and out of the workpiece. Figure 8.56
shows an example of a plain mandrel mounted between the centres of a lathe.

Fig. 8.56 Plain mandrels mounted between centres


Design of Fixtures 569

Expanding mandrels are more satisfactory for production work since the workpiece in
most cases is easier to remove and there is less wear on the mandrel as well as less
damage to the bore of the workpieces. There are many designs which may be used
by the tool designer, but before he spends a lot of time designing a special mandrel,
he should investigate the possibility of using one of the many commercial mandrels
available. It is often much less expensive to purchase a commercially made mandrel
than to design and build a special one. There are many special commercial mandrels
available, ranging from those with a self-contained hydraulic system which causes the
mandrel to expand to those containing a roller-actuated expanding sleeve.
A few mandrel designs which have been
used are shown in Figs 8.57 to 8.61. In
Fig. 8.57, a sleeve-type expanding man-
drel is shown. A sharp blow on the large
end of the mandrel with a lead hammer
causes the sleeve to expand inside the
workpiece. This type has an accuracy of
5 to 25mm, depending upon the care in
manufacture, and is commonly used as Fig. 8.57 Sleeve-type expanding mandrel
a general-purpose mandrel.
Figure 8.58 shows a simple expanding mandrel which can be quickly made for hold-
ing small parts. The design is for use in a collet lathe but could be changed to be held
in a collet or chuck quite easily. Such a mandrel is suitable only for small parts and
light lathe cuts, since the extreme ends of the mandrel may deflect when heavy pres-
sure is applied to the socket-head setscrew.

Fig. 8.58 Expanding mandrel designed to fit in place of standard collet

The mandrels in Figs 8.59 and 8.60 provide more holding power because the tapered
head acts on the extreme ends of the split and does not allow deflection. During
manufacture of this type, the split should be expanded a few thousandths and then
machined 0.025 mm larger than the bore of the part. When the draw is loosened, the
mandrel will collapse to allow its insertion into the part.
Sometimes, workpieces do not have enough bore surface to prevent slipping when
held on a mandrel. Figure 8.61 shows a mandrel equipped with an expanding shoe
which acts as a positive drive for the workpiece. In the case of a small bore in a large-
diameter workpiece, it may be necessary to provide an external device farther out on
the radius of the part in order to prevent slipping.
570 Tool Design

Fig. 8.59 Expanding mandrel designed to be held in standard collet or chuck

Fig. 8.60 Expanding collet mandrel

Fig. 8.61 Arbor equipped with expanding shoe

At times it is necessary to design a mandrel for holding previously threaded work-


pieces. Split sleeves on expanding mandrels can be expanded oversize, turned to the
major diameter of the thread, and then threaded to fit the internal thread of the part.
Another method sometimes used is to turn a shoulder on a plain mandrel and cut an
external thread up to the shoulder. The part is then screwed on the mandrel and the
turning operation performed. Such a mandrel may work well where very light cuts are
taken, but if heavy cuts are necessary, the workpiece will tighten against the shoulder
to the extent that it may take a pipe wrench to remove it.
The knock-off mandrel shown in Fig. 8.62 makes it easy to remove a threaded work-
piece. Note that the thread for the knock-off collar is left-handed, which resists any
Design of Fixtures 571

tendency to unscrew during the turning operation. The left-hand thread should be a
coarser pitch than that of the work itself, since the wedging action is much less in the
coarser-pitch thread and consequently releasing is easier. The work is released by a
sharp blow with a lead hammer on either of the two knock-off lugs.

Fig. 8.62 Mandrel for threaded parts

8.8 GRINDING FIXTURES


The design of grinding fixtures is basically the same as the design of fixtures for use
on other machine tools. For instance, there is little difference in a mill fixture and a
fixture to be used on a surface grinder likewise, a lather fixture is similar to a fixture
designed for use on a cylindrical grinder. There are, however, conditions unique to the
grinding process that must be taken into consideration.
The grinding process is usually a finishing operation where the workpiece dimensions
are held to close tolerances; therefore great accuracy is required in the design and
construction of grinding fixtures. Workpieces cannot deflect under clamping pressure,
and locating methods must be precise and exacting. Location of workpieces is made
easier, however, since by the time they are ready for grinding, most wokrpieces have
sufficient machined surfaces by which they can be positioned and held.
Intense, localised heat generated at the point of grinding-wheel contact is controlled
by a heavy application of coolant. Coolant nozzles must be designed properly and
located where they will do the most good. The entire contact zone between the wheel
and work should be flooded with an abundant supply of clean coolant. It should be
applied in large quantities sunder very little pressure. The nozzle should be designed
to offer the least possible restriction to the flow of the fluid. Surface grinding opera-
tions may call for two coolant nozzles to cool the workpiece efficiently during both the
forward and return strokes of the table. It may be necessary to provide a spray guard
to prevent splashing the operator and losing coolant.
Thin-walled tubing is difficult to grind because the thin wall does not dissipate the
heat produced by grinding and may cause burning. It is often possible to deliver the
572 Tool Design

coolant through the fixture and completely fill the inside of the tube with coolant. Since
the coolant on the inside of the tube dissipates the heat quickly, heavier cuts can be
taken, greatly reducing the grinding time required.
Grinding time can also be reduced if the wheel dresser is mounted directly on the
fixture. This practise eliminates the dead time required to move the wheel dresser to
the grinding wheel.

Magnetic chucking devices The standard method of holding parts for surface
grinding operations is magnetic chucking. Workpieces can be quickly mounted and
removed, and distortion caused by mechanical clamping is eliminated. A number of
parts can be mounted at one time, the only restriction being the size of the chuck.
The usefulness of the magnetic chuck can be extended by adapting fixtures directly
to the chuck surface. The magnetism will hold both the fixture and workpiece in place.
The basic fixture is made of a number of mild-steel plates separated from each other
by a nonferrous spacer. They are held together by nonferrous drawrods extending the
entire length of the fixture. The magnetic flux passes from the magnetic chuck through
the steel plates and the workpiece. The work is held securely to the plates and the
plates to the working surface of the chuck. This type of fixture is often known as a
magnetic-chuck parallel (see Fig. 8.63). Magnetic-chuck parallels can be machined to
accommodate various workpiece shapes, as shown in Figs 8.64 and 8.65.

Fig. 8.63 Use of magnetic-chuck parallels

Fig. 8.64 Use of magnetic-chuck V blocks


Design of Fixtures 573

Fig. 8.65 Use of magnetic-chuck fixtures

Odd-shaped workpieces are sometimes held on magnetic chucks with epoxy tooling
materials. The workpieces are coated with a releasing agent and mounted on the
work surface of the magnetic chuck. Liquid epoxy resin is poured around the workpe-
ices and allowed to harden. The workpieces can then be removed, leaving a master
impression in the solid epoxy. Additional workpieces can be placed in the impressions
and be held there by the magnetic force of the chuck. The solid epoxy prevents side
slippage and tilting of the part (see Fig. 8.66) epoxy tooling materials are described
in detail in Chapter 11.

Fig. 8.66 Use of epoxy tooling materials for fixture on magnetic chuck

Solved Design Problems

Problem 1 Design a lathe boring fixture for holding the workpiece for boring the
30 mm hole as shown in Fig. 8.67 (a) and (b). Consider that the four holes of 16 mm
diameters are accurately finished.

Solution: The given problem requires designing a lathe boring fixture to hold a par-
ticular workpiece on a lathe machine to carry out boring operation. The shape of the
workpiece is such that it will be difficult to hold and centre using standard fixture on a
lathe spindle. Therefore, there is a requirement of a special fixture to hold the work-
piece on a lathe spindle.
574 Tool Design

Fig. 8.67 The finished workpiece (a) Solid model (b) Detailed drawing

The designed parts of the strap clamp are presented in the Table 8.1 with reference to
the solid model and detailed drawing of individual parts.

Table 8.1 Lathe boring fixture parts for workpiece shown in Fig. 8.67
SI. no. Part name/Part description Quantity Figure Reference

1. Base Plate 1 8.68 (a) and (b)


2. Diamond Pin 1 8.69 (a) and (b)
3. Round Pin 1 8.70 (a) and (b)
4. Swing Clamp 2 8.71 (a) and (b)
5. Socket Head Screws 2 8.72 (a) and (b)
6. Faceplate 1 8.73 (a) and (b)
7. Hex bolts 4 8.74 (a) and (b)
8. Washer 4 8.75 (a) and (b)
9. Hex Nut 4 8.76 (a) and (b)

The fixture is designed using a standard lathe faceplate (refer to Fig. 8.73) along
with specially designed base plate (refer to Fig. 8.68). The workpiece is located and
clamped on the base plate (refer to Fig 8.78 and 8.79). The base plate along with the
workpiece is bolted on to the faceplate (refer to Fig. 8.79). To mount the assembly,
the threaded bore of the faceplate can be screwed on the lathe machine spindle. The
complete assembly and the exploded view of the designed fixture are shown in Fig.
8.79 and 8.80. The individual parts of the assembly are indicated in the exploded view
of the assembly in Fig. 8.80.
Design of Fixtures 575

Fig. 8.68 Base plate (a) Solid model (b) Detailed drawing

Fig. 8.69 Diamond pin (a) Solid model (b) Detailed drawing

Method of location Workpieces can be conveniently located by per machined


(drilled or bored) holed using two locating pins. In this case, two finished 16 mm
holes on opposite corners of the workpiece are used for location. Two locating pins
are used diagonally for the purpose of location (refer to Fig. 8.77). One of the locat-
ing pin is a round pin. A diamond pin is used as a secondary locator. Diamond pin is
used to reduce angular error, to reduce the jamming tendency and to allow the linear
576 Tool Design

variations of the workpiece. The design/selection of the diamond pin can made based
on the following calculations:
The diameter of the diamond pin, d is given as
___________
D2

where
÷
d = 2 ___ –WV – v2
4

D = diameter of hole to be located = 16 mm (in this case)


W = width of diamond pin = 4 mm (consider)
V = axial error in diamond pin location = 0.15 mm (consider)
Then, the diameter of the diamond pin, d = 15.92 mm

Fig. 8.70 Round pin (a) Solid model (b) Detailed drawing

Fig. 8.71 Swing clamp (a) Solid model (b) Detailed drawing
Design of Fixtures 577

Fig. 8.72 Socket head screw (a) Solid model (b) Detailed drawing

Fig. 8.73 Faceplate (a) Solid model (b) Detailed drawing


578 Tool Design

Fig. 8.74 Hex bolt (a) Solid modle (b) Detailed drawing

Fig. 8.75 Hex nut (a) Solid model (b) Detailed drawing

Fig. 8.76 Washer (a) Solid model (b) Detailed drawing


Design of Fixtures 579

Method of clamping and unclamping The workpiece is clamped on the base plate
using two swing clamps. The workpiece is first properly located on the base plate.
The clamping surface of the swing clamps are brought over the workpiece surface
and tightened over it using socket head screws. The socket head screws are screwed
on the base plate for tightening (refer to Fig. 8.79). For unclamping the socket head
screws are first loosened and both the swing clamps are swung away form the work-
piece surface. The workpiece is now free to be removed through the vertical axis.

Fig. 8.77 Assembly of faceplate, base plate and locating pins

Fig. 8.78 Assembly of faceplate, base plate, Fig. 8.79 Complete assembly of the
locating pins and the workpiece designed lathe boring fixture

Fig. 8.80 Exploded view of the designed lathe boring fixture


580 Tool Design

Problem 2 Design a lathe boring fixture for holding the workpiece for boring the 45
mm hole as shown in Fig. 8.15 (a) and (b).

Fig. 8.81 The finished workpiece (a) Solid model (b) Detailed drawing

Solution: The shape of the workpiece is such that it will be difficult to hold and centre
using standard fixture on lathe spindle. Therefore, there is a requirement of a special
fixture to hold the workpiece on a lathe spindle.
The designed parts of the strap clamp are presented in the Table 8.2 with reference
to the solid model and detailed drawing of individual parts.

Table 8.2 Lathe boring fixture parts for workpiece shown in Fig. 8.81
SI. no. Part name/Part description Quantity Figure Reference

1. Angle Plate 1 8.82 (a) and (b)


2. Swing Clamp 2 8.71 (a) and (b)
3. Socket Head Screws 2 8.72 (a) and (b)
4. Faceplate 1 8.73 (a) and (b)
5. Hex Bolt 1 8.74 (a) and (b)
6. Hex Bolt (long) 1 8.83 (a) and (b)
7. Counter weight 1 8.84 (a) and (b)
8. Washer 2 8.75 (a) and (b)
9. Hex Nut 2 8.76 (a) and (b)

The fixture is designed using a standard lathe faceplate (refer to Fig. 8.73) along
with specially designed angle plate (refer to Fig. 8.82). The workpiece is located and
clamped on the angle plate. The angle plate along with the workpiece is bolted on the
faceplate (refer to Fig. 8.85). To mount the assembly, the threaded bore of the face-
plate can be screwed on the lathe machine spindle. To the balance the rotating mass
of the workpiece and the angle plate, additional weight is attached to faceplate.
The complete assembly and the exploded view of the designed fixture are shown in
Fig 8.85 and 8.86. The individual parts of the assembly are indicated in the exploded
view of the assembly in Fig. 8.86.
Design of Fixtures 581

Fig. 8.82 Angle plate (a) Solid model (b) Detailed drawing

Fig. 8.83 Hexbolt long (a) Solid model (b) Detailed drawing
582 Tool Design

Fig. 8.84 Counter weight (a) Solid model (b) Detailed drawing

Method of location Workpieces in this case have tapered side surface, making it
difficult to locate using locating pins. Therefore, in this case, a cavity on the angle
plate, of the shape of the outer profile of the workpiece is used for location (refer to
Fig. 8.82). The workpiece can be slid into the cavity of the angle plate for location.

Method of clamping and unclamping The workpiece is clamped on the angle plate
using two swing clamps. The workpiece is first properly located on the angle plate.
The clamping surface of the swing clamps are brought over the workpiece surface
and tightened over it using socket head screws. The socket head screws are screwed
on the angle plate for clamping (refer to Fig. 8.85).
For unclamping, the socket head screws are first loosened and both the swing clamps
are swung away for the workpiece surface. The workpiece is now free to be slid out
of the cavity locator.

Fig. 8.85 Complete assembly of the designed lathe boring fixture


Design of Fixtures 583

Fig. 8.86 Exploded view of the designed lathe boring fixture

Summary

Production jobs often require accurate locating, and holding of the workpiece. This
is achieved using fixtures. Unlike jigs, fixture does not guide the tool. In construction,
the fixtures comprise different standard or specially designed work holding devices,
which are clamped on the machine table to hold the work in proper location. This
chapter deals with principles, methods and devices used for fixturing.
• Fixture holds and positions the work but does not guide the tool.
• Fixtures are generally heavier in construction and are bolted rigidly on the
machine table, whereas the jigs are made lighter for quicker handling, and
clamping with the table is often unnecessary.
• Fixtures are employed for holding work in milling, grinding, planing, or turning
operations, whereas the jigs are used for holding the work and guiding the tool
particularly in drilling, reaming, or tapping operations.
• Depending on the application, fixtures may of various type; Milling fixture,
Lathe fixture, Boring fixture, Broaching fixture, Grinding fixture, etc.
• Workpiece which cannot be held by a standard fixture (e.g., vise, chuck, man-
drels, collets etc.), requires specially designed and fabricated fixture.
• A mill fixture holds the part in the correct relation to the milling cutter as the table
movement carries the part through the cutters. It consists of five main parts,
the base, clamps, rest blocks or nest locating points, and gauging surfaces.
• Fixtures for broaching are often simple plate, plug, or horn fixtures which
established the correct position of the workpieces in relation to the broach and
machine faceplate.
• A large majority of lathe operations can be accomplished by using standard
chucks and holding methods. However, many parts such as castings and forg-
ings cannot readily be mounted on any of the standard fixtures. It is therefore
necessary to manufacture special work-holding fixtures for machining these
parts.
584 Tool Design

Questions
1. How was the term fixture derived?
2. What is the major difference between a drill jig and a fixture?
3. What are the economic aspects of the use of a fixture?
4. Why are keys mounted to the base of mill fixtures?
5. Why must the clamps on milling fixtures be extremely rigid?
6. How can the accuracy of contact between bearing surfaces of the mill fixture
and the workpiece resting upon it be checked?
7. What is the common setting distance used on mill-fixture setting gauges?
8. What are the various methods of production milling?
9. What are the major advantages of magnetic and vacuum milling fixtures?
10. What is considered the minimum area of a workpiece that is held by a
vacuum?
11. What are the two general classes of boring fixture?
12. How does clamping over unsupported surfaces produce nonparallel holes in
bored workpieces?
13. Why is it necessary for the rear pilot of a double-piloted boring bar to be larger
than the sweep of the cutter?
14. How are broaching fixtures classed?
15. What precautions should be taken when designing external broaching
fixtures?
16. What is the basic difference between a lathe fixture and the fixture previously
discussed?
17. Referring to Question 16, what additional considerations apply to lathe
fixtures?
18. What is the rule of thumb for the correct gripping area of chuck jaws?
19. What is the advantage of using a rocking jaw on a three-jaw chuck?
20. What is the advantage of wraparound chuck jaws?
21. What is the result if the bore diameter of wraparound jaws is not held within ±
0.025 mm?
22. How is a faceplate fixture located on the lathe faceplate?
23. What is the advantage of magnetic and vacuum chucks?
24. What is the standard taper on a plain mandrel?
25. What conditions unique to a grinding process must be taken into consideration
when designing grinding fixtures?
26. Why is the wheel dressing often attached directly to the grinding fixture?
27. What is the standard method of holding parts for surface-grinding operations?
Design of Fixtures 585

Design Problems

Each student will be assigned one of the following fixture design problems. Complete
a finished tool drawing using the tool-design procedure explained in Chap. 1. The
student will be required to determine the type of machine tool best suited for the
machining operation. Use an existing machine tool in the school shop or determine
machine-tool dimensions from manufactures’ catalogues.
1. Design a milling fixture to machine the end surfaces and flange edges of the
link connecting rod in Fig. 8.87a. Assume that the end holes have been previ-
ously bored to size.
2. Design a milling fixture to end-mill the channels of the link connecting rod in
Fig. 8.87a. Assume that the end holes have been previously bored to size.
3. Design a milling fixture to mill the flange flank of the link connecting rod in Fig.
8.87a. Assume that the end holes have been previously bored to size.
4. Design a milling fixture to mill the sides of the large and small ends of the con-
necting rod in Fig. 8.87b. Assume that the end hole has been previously bored
to size.

Fig. 8.87 Milling-fixture design problems: (a) link connecting rod (b) connecting rod (surface-
milled chrome-molybdenum forging)
586 Tool Design

5. Design a milling fixture to mill a flat on opposite sides of the cylinder-barrel


flanges on the airplane cylinder in Fig. 8.88. Assume that the workpiece has
been previously bored and faced. All holes in the flange are also drilled.
6. Design a milling fixture to mill the finished surface of the casting in Fig. 8.89a.
This is the first machining operation on the rough casting. Holes are drilled
after the surface is milled.
7. Design a milling fixture to mill the finished surface on the clevis in Fig. 8.89b.
Holes will be drilled after the surfaces have been machined.
8. Design a milling fixture to mill the 50 mm slot shown in the right-angle clevis
in Fig. 8.89c. Assume that the hole and 27 mm slot have been previously
machined.
9. Design a boring fixture to bore the holes in the connecting-rod bushings in Fig.
8.87a. Bushings are to be bored after they are pressed into place. The bush-
ings are not shown in the drawing.
10. Design a lathe fixture to turn the OD of the pulley in Fig. 8.90a. Assume 16 mm
hole has been previously machined.
11. Design a lathe fixture to turn the 12.5 mm shank of the oarlock in Fig. 8.90b.
12. Design a broaching fixture to broach the slots marked with finishing symbols
on the spring chair in Fig. 8.91a.
13. Design a broaching fixture to broach the hole in the double-hexagon box
wrench in Fig. 8.91b.
14. Design a broaching fixture to broach the slots in the V-block clamp in Fig.
8.92.
15. Design a grinding fixture to surface grind the wokrpiece in Fig. 8.89d. work-
piece is 1090 heat-treated steel, and 0.4 mm per side must be removed.
Design of Fixtures 587

Fig. 8.88 Airplane cylinder


588 Tool Design

Fig. 8.89 Milling-fixture design problems

Fig. 8.90 Lathe-fixture design problems


Design of Fixtures 589

Fig. 8.91 Broaching-fixture design problems: (a) Spring chair (b) Double-hexagon box
wrench
590 Tool Design

Fig. 8.92 V-block clamp

References
Books
• ASM: “Matals Handbook,” 8th ed., “Matching,” Metals Park, Ohio, 1967.
• ASTME: “Handbook of Fixture Design,” FW Wilson (ed.), McGraw-Hill, New york, 1962.
• “Tool Engineers Handbook,” 2d ed., FW Wilson (ed.), McGraw-Hill, New York, 1959.
• Cincinnati Milacron, Inc.: “A Treatise on Milling Machines,” 3d ed., Cincinnati, Ohio, 1951.
• Grant, HE: “Jigs and Fixtures,” McGraw-Hill, New York, 1967.
• Habricht, FH: “Modern Machine Tools,” Van Nostrand, Princeton, NJ, 1963.
Catalogues
• Accurate Bushing Co., Garwood, NJ
• American Drill Bushing Co., Los Angeles.
• Carr Lane Manufacturing Co., St Louis, Mo.
• Jergens, Inc., Cleaveland, Ohio.
• Universal Engineering Co., Frankenmuth, Mich.
• Vlier Engineering Corp., Los Angeles.
DESIGN OF SHEET-METAL
BLANKING AND PIERCING 9
DIES

9.1 INTRODUCTION
The casual observer seldom takes a second look at the punch-press department
when he visits a modern production plant. He views this section of the plant as an
assemblage of noisy mechanical monsters clamly out parts from a roll of strip metal
and is much more concerned with newer matching and manufacturing processes
often referred to as automation.
Numerical control is a prime example. It has been called the major factor in auto-
mating small-lot production, and probably more has been written on numerical con-
trol in technical journals than any other manufacturing process since its introduction.
The whole community turns out when a local company opens its doors to show its
new numerically controlled machining centre. Those associatad with metal working
and manufacturing must at least have a talking knowledge of numerical control to
be counted as members of the group. Yet, with all its popularity and contributions to
modern manufacturing, numerical control can probably measure about a mm on the
yardstick used to measure the contributions of the lowly punch press. Hardly any item
in the laundry or kitchen cannot credit its existence to the punch press. The average
person could not afford an automobile if it were not for the punch press or one of its
relatives. The 25-cent can opener would not exist without the technology of stamped
metal parts.
The causal observer should stop and remember that a completed workpiece is pro-
duced at each noisy “chomp” of the press ram. What is more amazing is that sheet-
metal presswork methods date back longer than one can imagine when speaking of
automated processes. The technology of sheet-metal presswork emerged with the
development of the steel industry, and to a large degree we owe our present standard
of living to the production of stamped metal parts. The numerically controlled, but the
punch press and its relatives in the far end of the plant are still our bread and butter.
This chapter is intended to acquaint the student or beginning tool designer with sheet-
metal blanking, piercing, and related operations. This design of blanking and piercing
dies will be discussed in some detail.
It is still an industrial practise to specify presses ratings and related press working
parts in tons and inches. Therefore, in tune with industrial practise, the units of this
chapter are mostly in tons and inches.

9.2 INTRODUCTION TO DIE-CUTTING


OPERATIONS
9.2.1 The Fundamentals of Die-Cutting Operations
While there are many die-cutting operations, some of which are very complex, they
can all be reduced to the following simple fundamentals:
592 Tool Design

Plain blanking Figure 9.1 shows a simple operation of this type. The material
used is called the stock and is generally a ferrous or nonferrous strip. During the
working stroke, the punch goes through the material, and on the return stroke, the
material is lifted with the punch and is removed by the stop pin is a gauge for the
operator. In practise, he feeds the stock by hand and locates the holes to be punched
as shown. The part that is removed from the strip is always the workpiece (blank)
in a blanking operation. Subsequent pressworking operation may be performed on
the blank. Figure 9.2 shows a key made in three punch-press operations, the first of
which is blanking.

Fig. 9.1 Plain Blanking

Piercing This operation consists of simple hole punching.It differs from blanking in
that the punching (or material cut from stock) is the scrap and the strip is the workpiece.
Piercing is nearly always accompanied by a blanking operation either before, after, or
at the same time. Figure. 9.3a shows a typical blanked and pierced workpiece.

Lancing This is a combined bending and cutting operation along a line in the work
material. No metal is cut free during a lancing operation. The punch is designed to cut
on two or three sides and bend along the fourth side. Figure 9.3b shows the principle
of the lancing operation.

Cutting off and parting A cut-off operation separates the work material along a
straight line a single-line cut. When the operation separates the work material along
a straight-line cut in a double-line cut, it is known as parting. Cutting off and parting
operations are used to separate the scrap strip. Cutting off and parting usually occur
in the final stages of a progressive die.
Cutting off is also used to chop up the scrap strip skeleton as it leaves the die. This
makes the scrap much easier to handle. Figure 9.3c and d show the basic principles
of cutting off and parting.
Design of Sheet-Metal Blanking and Piercing Dies 593

Fig. 9.2 Blanked and coined workpiece (Revere Copper & Bruss, Inc.)

Fig. 9.3 Basic die-cutting operations: (a) Blanked part pierced holes (b) Lancing (c) Cut off
(d) Parting (e) Notching (f) Shaving
594 Tool Design

Notching This operation removes metal from either or both edges of the strip.
Notching serves to shape the outer contours of the workpiece in a progressive die or
to remove excess metal before a drawing or forming operation in a progressive die.
The removal of excess metal allows the flow or from without interference from excess
metal on the sides. Figure 9.3e shows a typical example of notching.

Shaving Shaving is a secondary operation, usually following punching, in which the


surface of the previously cut edge is finished smoothly to accurate dimensions. The
excess metal is removed much as a chip is formed with a metal-cutting tool. There is
very little clearance (close to zero) between the punch and die, and only a thin selec-
tion of the edge is removed from the edge of the workpiece. Figure 9.3f describes the
shaving operation.

Trimming This operation removes the destorted excess metal from drawn
shapes and also provides a smooth edge. Trimming is discussed more in detail in
Chapter 10.

9.3 POWER-PRESS TYPES


The types of power presses available for metal-cutting and forming operation are
varied, the selection depending upon the type of operation. Not all types of presses
will be described because of space limitations. The basic types of presses and press
mechanisms will be described to give the beginner the necessary background for
designing press tooling.
Presses are classified by (i) type of frame, (ii) source of power, (iii) method of actua-
tion of slides, (iv) numbers of slides incorporated, and (v) intended use. Most presses
are not classified by only category one but several. For example, a straight-side press
may be mechanically or hydraulically driven
and may be either single- or double-acting.

Classification by frame type The frame of


a press is fabricated by casting or by weld-
ing heavy steel plates. Cast frames are quit
stable and rigid but rigid but expensive.
Cast-frame construction also has the advan-
tage of placing a mass of material where it
is needed most. Welded frames are gener-
ally less expensive and are more resistant to
shock loading because of the greater tough-
ness of steel plate.
The general classification by frame includes
the gap frame and the gap frame and the
straight side. The gap frame is cut back
below the ram to from the shape of a letter
C. This allows feeding a strip feeding from
front to back or ejection of finished parts out
the back. Gap-frame presses are manufac-
Fig. 9.4 OBI press (Service Machine
tured with solid frames fixed in a vertical or
Company, Inc.)
inclined position. Others are manufactured
Design of Sheet-Metal Blanking and Piercing Dies 595

with a separate frame mounted in a base which allows the frame to be inclined at an
angle in three different positions.
The reason for inclining the press is to allow part to fall through the open back by
gravity. The three-position inclinable press is frequently referred as an open-back
inclinable (OBI) press (see Fig. 9.4). Solid gap-frame presses are obtainable in higher
tonnages than inclinable ones because of the rigid base and solid construction.
The OBI press is the most common press in use today. It ranges from a small 1-ton
bench press to floor presses rated up to 150 tons. Its main use is for blanking and
piercing operations on relatively small workpiece, although bending, forming, and
drawing operations can also be done.
Figure 9.5 shows the major components of an
OBI press, as follows:
1. A rectangular bed, the stationary and
usually horizontal part of the press,
serving as a table to which a bolster
plate is mounted.
2. A bolster plate, consisting of a flat steel
plate from 50 to 125 mm thick, secured
to the press for locating and supporting
the die assembly.
3. The ram, sometimes called the slide,
which reciprocates within the press
frame and to which the punch or upper-
die assembly is fastened.
4. A knockout, consisting of a crossbar
through a slot in the ram that contacts a
pin in the die to eject the workpiece. Fig. 9.5 Major components of an
5. The flywheel, absorbs energy from OBI press
the motor continuously and delivers
its stored energy to the workpiece intermittently, making it possible to use a
smaller motor.
6. The pitman, consisting of a connecting rod to convey motion and pressure from
the main shaft or eccentric to the press slide.
The straight-side frame press incorporates a slide or ram which travels up and down
between two straight sides, or housings, and is commonly used for large and heavy
work. The size of the press is limited to some extent because the housings reduce
the working area. However the frame constriction does permit large bed, although
underdrive presses may be obtained with the drive mechanism located below the
bed. Straight-side presses are classified as single- two- or four-point suspension,
depending upon the number of connections between the slide and the main drive
shaft. Figure 9.6 shows a typical straight-side press.

Classification by source of power The great majority of presses receive their power
mechanically or hydraulically. A few manually operated presses are hand-operated
through levers or screws, but they are hardly suited for high production.
596 Tool Design

Mechanical presses use a flywheel-


driven system to obtain ram movement.
The heavy flywheel absorbs energy
from the motor continuously and deliv-
ers its stored energy to the workpiece
intermittently. The motor returns the
flywheel to operating speed between
strokes. The permissible slowdown of
the flywheel during the work period is
about 7 to 10 per cent in nongeared
presses and 10 to 20 per cent in geared
presses. The flywheel is attached
directly to the main shaft of the press
(nongeared), or, it is connected to the
main shaft by a gear train. Nongear
drives are used on presses of low ton-
nage and short strokes. The number of
stokes per minute on nongear drives is
usually quit high. Fig. 9.6 Single-action straight-side eccentric
shaft mechanical press (Varson All-
Gear-driven presses transmit the steel Press Company)
energy of the flywheel through a single
or double-reduction gear. The single-reduction gear drives is suited for heavy blank-
ing operations or shallow drawing. The double-gear drive is used on large, heavy
presses where it is necessary to move large amount of mass at slower speeds. The
double reduction greatly reduces the strokes per minute without reducing the flywheel
speed. Figure 9.7 shows the basic types of mechanical drives.
Hydraulic presses have a large cylinder and piston, coupled to a hydraulic pump. The
piston and press ram are one unit. The tonnage capacity depends upon the cross-
sectional area of the piston (or pistons) and the pressure developed by the pump. The
cylinder is double-acting in order to move the ram in either direction. The advantage
of a hydraulic press is that it can exert its full tonnage at any position of the limits of
the hydraulic-cylinder travel. The speed and pressure are also constant throughout
the entire stroke. Figure 9.8 shows a typical hydraulically driven press.

Classification by method of slide actuation The flywheel of a press drives the main
shaft, which in turn changes the changes the rotary motion of the flywheel into the
linear motion of the slide or ram. This is generally accomplished by incorporating
crankpins or eccentrics into the main drive shaft as shown in Fig. 9.7. The number of
throws or eccentric on the main shaft is determined by the number of points of sus-
pension of the slides. Points of connections, called connecting rods or pitmans, are
usually adjustable in length so that the shut height of the press can be varied.
The most common driving device is the crankshaft, althought many newer presses
use the eccentric for ram movement. The main advantage of the eccentric is that it
offers more surface area for bearing support for the pitman, and main disadvantage is
its limitations on the length of stroke. Therefore, presses having longer strokes gener-
ally utilise the crankshaft.
Design of Sheet-Metal Blanking and Piercing Dies 597

Fig. 9.7 Basic types of mechanical press drives: (a) Nongeared (b) Single-reduction gear
and (c) Double-reduction gear drive

In addition to eccentric and crankpins, slides


are actuated by cams, toggles, rack and
pinions, screws, and kunckles. Space does
not permit discussion of these mechanisms
in this text. Information may be obtained by
referring to the various die-design and press
handbooks.

Classification by the number of slides incor-


porated The number of slides incorporated
in a single press is called the action, i.e., the
number of rams or slides on the press. Thus
a single-action press has one slide. A double-
action press has two slides, an inner and an Fig. 9.8 Typical hydraulic press (Ver-
outer slide carries the blank holder and the son Allsteel Press Company)
598 Tool Design

inner slide carries the punch. The outer, or blank while than the inner, or punch-holder
slide, dwells to hold the blank while the inner slide continues to descend, carrying the
draw punch to perform the drawing operation.
A triple-action press is the same as a double-action with the addition of a third ram,
located in the press bed, which moves upward in the bed soon after the other two
rams descend. All three slide movements are properly synchronised for triple-action
drawing, redrawing, and forming.

9.4 GENERAL PRESS INFORMATION


The tool designer must know certain fundamentals of press operation before he can
successfully design press tooling.

Press tonnage The tonnage of a pressure to the force that the press ram is able to
exert safely. Press slides exert forces greater than the rated tonnage because of the
built-in safety factor, but this is not a license to overload.
The tonnage of hydraulical presses is the piston area multiplied by the oil pressure in
the cylinder. The tonnage is varied by changing the oil pressure.
The tonnage of mechanical presses is determined by the size of the bearings for
the crankshaft or eccentric and is approximately equal to the shear strength of the
crankshaft metal multiplied by the area of the crankshaft bearings. The tonnage of
a mechanical press is always given when the slide is near the bottom of its stroke
because it is greatest at this point.

Stroke The stroke of press is the reciprocating motion of a slide, usually specified
as the number of the number of the motion. The stroke is constant on a mechanical
press but adjustable on a hydraulic press.

Shut height The shut height of a press is the distance from the top of the bed to the
bottom of the slide with the shut height of the press.
The shut height of a press is always given with the adjustment up. Lowering the
adjustment of the slide may decrease the opening of a press from the shut height
down, but it does not increase the shut height. Thus, the shut height of a die must not
be greater than the shut height of the press. It may be less, because the die opening
in the press can be reduced by lowering the adjustment.

Die space Die space is the area available for mounting dies in the press.

9.5 MATERIAL-HANDLING EQUIPMENT


Press feeding may be done manually. For example, the simple die shown in Fig. 9.1 is
designed for manual feeding. Here a stop pin is built into the die block to catch in the
hole resulting from the blanking operation. When the punch moves up on the return
stroke, the strip is raised enough to clear the pin and allow the strip to advance to the
next position. The operator needs only apply pressure on the strip in direction of feed.
The strip is generally short for easy handling.
Design of Sheet-Metal Blanking and Piercing Dies 599

Manual feeding is satisfactory for low production or with presses having a low stroke-
per-minute capability, but modern presses as a rule operate at 200 to 300 strokes per
minute, with exceptions running at these speeds.
Automatic feeding devices are the answer to high-speed press production. They
enable steel mills to prepare strip in large coils especially for use with punch presses.
When the coils reach the punch press, they are fed continuously through the die
with a resulting increase in production. Heavier-gauge strip can also be used when
the press utilises devices. The following descriptions cover only a few of the feeding
devices commercially available, but they will suffice to give a basic understanding of
this equipment.

Coil-unwinding equipment The two basic types of coil-unwinding equipment are


the reel and coil cradles. Of the two, the reel is considered the best all-around method
of playing off stock since it does not scratch or mar the strip in any way. It may
or may not be power- driven. Power-driven models contain a long loop arm with a
roller attached that rides on the uncoiling sheet stock. When enough stock has been
uncoiled, the loop arm actuates a switch that stops the power driven. As the stock
is used up, the loop arm is raised and the power drive starts the reel uncoiling more
metal. The reel always keeps the correct amount of stock unwound in the form of a
loop. The press-feeding mechanism draws stock from the loop. This prevents slipping
and marring of the stock and marring of the stock in the feeding mechanism.
Unpowered reels require the stock to be uncoiled by an external power source, which
may be the feeding mechanism or straightening rolls (described later). Unpowered
reels are equipped with an automatic or manual brake that prevents the reel from
turning after sufficient metal has been uncoiled. The manual brake has a predeter-
mined drag which can be varied with an adjusting nut. On the automatic braking reel,
a loop arm control permits stock to be played out with minimum drag and friction.
When enough stock has been uncoiled, a loop will form and the arm will drop down,
causing the brake to be applied. Figure 9.9 shows a powered reel and Fig. 9.10
shows an unpowered coil reel.

Fig. 9.9 Motor-driven coil-unwinding reel Fig. 9.10 Unpowered plain coil-unwinding
(F. J. Littell Machine Co.) reel (F J Littell Machine Co.)
600 Tool Design

The cradle supports the stock on the outside diameter of the coil. The coil nests
against rollers or a small conveyor built into the base of the cradle. The cradle can
scratch the stock surface because the cradle rolls contacts the stock being uncoiled.
The cradle rolls are generally power- driven. Cradles are generally less expensive
than equivalents-size reels. A cradle is shown in Fig. 9.11.

Fig. 9.11 Using a coil cradle in operation with a roll feed (F J Littell Machine Co.)

Strip-straightening devices The primary function of a straightening machine is to


remove wrinkles and curvature after uncoiling. Strip straightening is an intermediate
step between uncoiling and feeding into the press. The straightening machine con-
sists of a series of interacting rolls which bend the stock back and forth past its elastic
limits (see Fig. 9.12). The first roll gives the greatest bend, the second somewhat less,
and so until the last roll, which does very little work. Five rolls are generally all that
are required to remove coil set, seven to eleven rolls being needed to flatten the stock
generally For very thin stock, seventeen rolls may be required. Thin and hard stock
generally requires smaller diameter rolls, close spacing, and more rolls. Heavier stock
require larger diameter and greater spacing.
Stock straighteners may be mounted on the press, in a separate machine, or in com-
bination with a stock reel or coil cradle.
For economic reasons, straightening rolls are generally combined with feed rolls
or with a cradle. Figure 9.13 shows a combination straightening machine and coil
cradle.

Fig. 9.12 Principle of straightening devices


Design of Sheet-Metal Blanking and Piercing Dies 601

Strip-feeding equipment After uncoil-


ing and straightening, the final step of
stock handling is feeding the strip into the
press, the most exacting and critical step
in the process. Numerous strip-feeding
devices are commercially available, and
the selection depends upon such factors
as width, thickness, and surface condi-
tion of the material, press speed, feeding
length, and feeding speed-to mention a
few. Fig. 9.13 A combination straightening
machine and coil cradle (F J
The two main categories of feeding sys- Littell Machine Co.)
tems ate the roll feed and the slide (or
hitch) feed. The roll feed moves the stock between rollers; the slide feed grips the
stock and slides or hitches it forward intermittently.
A roll feed consists essentially of a pair of rolls driven through an overrunning clutch
or ratchet mechanism timed from the press main shaft or slide. Roll feeds may be
single or double. Single feeds are mounted on only one side of the press and can be
arranged to push or to pull the stock through the die set. They are generally used with
dies having no scrap. It this type of work, the stock is most commonly pushed into the
die set by the roll feed, and the scrap skeleton is cut into small pieces on the opposite
side of the die set. A single-roll feed is shown in Fig. 9.14 mounted on a gap-frame
press.
Double-roll feeds consist of a roll-feed mechanism mounted on each side of the press
bed with a drive connection between them. One feed pushed the stock in while the
other pulls it out to keep it tight and prevent buckling in the die set. No short length of
strip remain after the stock has been fed through the dies, as with a single-roll feed.
Fig. 9.15 shows a double roll feed.

Fig. 9.14 A single-roll feed mounted on Fig. 9.15 A double rack-and-point roll feed
an OBI press (FJ Littell Ma- equipped with straightener and
chine Co.) scrap cutter mounted on an OBI
press (F J Littell Machine Co.)
602 Tool Design

The combination feed and straightening machine shown in Fig. 9.16 would be classed
as an independent power-feed roll. The feed rolls in this machine are entirely inde-
pendent of the press slide movement and therefore some means of timing the rolls
must be provided. This type of feeding mechanism is used with larger presses and is
made to handle wider coils. The power required to feed this heavy material cannot be
obtained by direct to the press slide or crankshaft.

Fig. 9.16 Combination feeding and straightening machine (F J Littell Machine Co.)

Slide or hitch feeds use grippers that grasp the strip mechanically and carry it forward
into the die set by a reciprocating mechanism, driven from the press crankshaft a cam
mounted on the punch holder or shoe, hydraulically or pneumatically. The mecha-
nism basically consists of a feed block mounted on a slide or guide bar. Grippers are
mounted on the feed block and hold the strip firmly as the stock is moved forward. The
stock is released on the return stroke. A check mechanism prevents the strip from
slipping back while the gripper is released. The feed block travels between adjustable
stops, which determine the feed length. Figures 9.17 and 9.18 show typical hitch-
feeding mechanisms.
Air-slide feeds are advanced by a pneumatic cylinder. Their major advantages are
that they have longer feed lengths and need not be driven by press ram or crankshaft.
Timing is determined by an air value mounted on the press ram or crankshaft.

Fig. 9.17 Slide or hitch feeding mechanism (The Producto Machine Company)
Design of Sheet-Metal Blanking and Piercing Dies 603

Fig. 9.18 Hitch feeding mechanism (The Dickerman Company, Division of Standerd Interna-
tion Corp.)

9.6 CUTTING ACTION IN PUNCH AND


DIE OPERATIONS
The cutting action that occurs in blanking or piercing is quit similar to that of chip
formation ahead of a cutting tool. The punch contacts the work material supported
by the die and a pressure build-up occurs. When the elastic limit of the work material
is exceeded, the material begins to flow plastically (plastic deformation). The punch
penetrates the work material, and the blank, or slug, is displaced into the die open-
ing a corresponding amount. A radius is formed on the top edge of the hole and the
bottom edge of the slug, or blank, as shown in Fig. 9.19a. The radius is often referred
to as rollover, and its magnitude depends upon the ductility of the work material.
Compression of the slug material against the wall of the die opening burnishes a por-
tion of the edge of the blank as shown in Fig. 9.19b. At the same time, the plastic flow
pulls the material around the punch, causing a corresponding burnished area in the
work material. Further continuation of the punching pressure then starts fractures at
the cutting edge of the punch and die (see Fig. 9.19c). Under ideal cutting conditions,
the fractures will meet and the remaining portion of the slug edge will be broken away.
A slight tensile burr will be formed along the top edge of the slug and bottom edge of
the work material (see Fig. 9.19d).
Figure 9.20 shows the characteristic appearance of the edges of parts produced by
blanking and piercing operations in detail. The edge radius (or rollover) is produced
during the initial stage of plastic deformation. The edge radius is more pronounced
with soft materials.
604 Tool Design

Fig. 9.19 Cutting-action progression when blanking or piercing metal

Fig. 9.20 Characteristic appearance of edge of parts produced by piercing and blanking
Design of Sheet-Metal Blanking and Piercing Dies 605

The highly burnished band is the result of the material’s being forced against the walls
of the punch and die and rubbing during the final stages of plastic deformation. The
sum of the edge radius depth and the burnished depth is referred to as penetration,
i.e., the distance the punch penetrates into the work material before fracture occurs.
Penetration is usually expressed as a per cent of material thickness, and it depends
upon the properties of the work material. As the work material becomes harder, the
per cent of penetration decreases. For this reason, harder materials have less defor-
mation and burnished area.
The remaining portion of the cut is the fractured area or break. The angle of the frac-
tured area is the breakout angle. The tensile burr is adjacent of the break. The burr
side of a blank or slug is toward the punch, and the burr side of the work material is
toward the die opening.

9.7 DIE CLEARANCE


Clearance is defined as the intentional space between the punch cutting edge and
die cutting edge. Clearance is always expressed as the amount of clearance per side.
Theoretically, clearance is necessary to allow the fracture to meet when break occurs
as shown in Fig. 9.21. The amount of clearance depends upon the kind, thickness and
hardness of the work material.

Fig. 9.21 The effect of clearance: (a) Too little clearance: fracture do not meet (b) Correct
clearance: fracture meet

Excessive clearance allows a large edge radius (rollover) and excessive plastic defor-
mation. The edges of the material tend to be drawn or pulled in the direction of the
working force, and the break is not smooth. Large burrs are present at the break
edge.
Insufficient cutting clearance causes the fractures to miss and prevents a clean break,
as shown in Fig. 9.21a. A partial break occurs, and a secondary break connects the
original or main fractures. This is often referred to as secondary shear. The secondary
break does not allow separation of the material without interference, and a second
burnished ring is formed as shown in Table 9.1. The burnished ring may appear as
a slight step around the outside edge of the blank or around the inside edge of the
hole. Insufficient clearance increases pressure on the punch and die edge and has a
marked effect on die life.
Authorities disagree on the correct amount of clearance for a given workpiece mate-
rial. One reference may recommend six per cent of stock material thickness per side,
while another may recommend 12 per cent for the same material. The difference is
606 Tool Design

Table 9.1 Standard edges for stampings


Section Description and use Typical clearance.
Material % per side

This edge is obtained at the


upper limits of clearance for a
usable blank. The blank has a
large radius on the bottom, a
large angle on the side, and a
large tensile burr on he top 304 stainless 23
edge. This type is generally 1020 mild steel 21
used only on structural metal- 2-SO aluminium 17
working machines where stand-
ard punches and dies are used
to pierce a wide range of
material thicknesses and the
make purpose is to produce a
hole
This edge leas a large radius on
the bottom, a normal tensile
burr, and a lesser edge angle
than the edge above. This edge
will provide maximum de life 304 stainless 13
and a blank tat is assayable 1020 mild steel 12
quality for general work. The 2-SO aluminium 9
clearance range from a low of 1070 high-carbon
6% on stock thickness per since steel 18
for magnesium, to a high of
18% on high-carbon and alloy
steel.
This edge is considered the most
desirable for most purposes as
it is burr-free and has a normal
edg erdius on the bottom of
the blank and a shallow edge
angle. This edge is particularly
desirable for use in parts made
of work-hardenable material 304 stainless 10
which will undergo severe forming 1020 mild steel 7
operations. The clean stress-free 2-SO aluminium 7
edge reduces to a minimum the Annealed brass 6
possibility of edge cracking during
forming operations. The clearance
on this edge may vary from a low
of 4% of stock thickness per side
on magnesium to a high of 15%
per side on high-carbon steel.
This edge is considered very
desirable for stampings used for
mechanisms or parts that require
edge finishing such as sheaving
or polishing. The edge has a
Design of Sheet-Metal Blanking and Piercing Dies 607

Section Description and use Typical clearance.


Material % per side
minimum radius on the blank, 304 stainless 4
1
nearly perpendicular edges, and 1020 mild steel 6 __
2
a normal tensile burr. It can be 2-SO aluminium 3
1
recognised yb the spotty showing Annealed brass 2 __
2
of secondary shear on the break
of the blank. Clearnces on this
edge range from a low of 2% per
side to a high of 12% of stock
thickness per side.
This edge has a minimum radius,
a perpendicular edge, and a normal
tensile burr and is recognised by
the complete secondary shear on
the edge. This edge is normally
disastrous in terms of die life 2-SO aluminium 1
on harder materials (as low as Copper 1
1
four blanks from this type of Annealed brass 1 __
2
edge on 1070 material). However, 1020 mild steel 2
this edge can be useful on some
of the softer materials, such as
brass, lead, soft copper, and alum-
inium, with reasonable die life.

probably in the end result each is striving for the same material. The difference is
probably in the end result each is striving for. The designer should consider the appli-
cation of the pierced or blanked workpiece. When the purpose is only to make a hole,
as in the case of structural steel, wide clearances may be used to increase die life.
Blanked workpieces that assemble as an integral part of a mechanism require tighter
clearances. Table 9.1 shows various edges for stampings with a description and use
of each. Note that the recommended clearance varies from 2 to 21 per cent for 1020
mild steel. The graph in Fig. 9.22 shows standard die clearance for different edges in
12 common materials.
The diameter of the blank or pierced hole is determined by measurement of the bur-
nished area. Since the burnished area on the blank is produced by the walls of the
die, the diameter of the blank will be the same as the diameter of the die (disregarding
a slight expansion after the blank is pushed from the die). The same principle applies
to the diameter of the pierced hole. The burnished area in the hole is caused by the
punch; thus the diameter of the pierced hole will be the same as the punch. Therefore,
die clearance is either placed or the die, depending upon whether the pierced hole
or the blank will be the desired workpiece. If the blank is to become the workpiece,
the die diameter is made to the workpiece size and the punch is reduced is size to
an amount equal to the die clearance. If the pierced hole is to become part of the
workpiece, the punch is made to the correct hole size and the die opening is made
oversize to an amount equal to the die clearance. In simple terms, the die controls
blank size and the punch controls hole size.
608 Tool Design

Fig. 9.22 Standard die clearances in per cent of stock material thickness per side (From
Machine and Tool Bluebook, by permission)

Angular clearance During the metal-


cutting operation some of the grain
structure is stressed to a point below
the elastic limit. When the cutting
operation is completed a small amount
of springback occurs, causing the
remaing hole in the stock material to
reduce in diameter and grip tightly on
the punch. The blanked part increases
in diameter and grips tightly inside the
die opening. A small amount of draft is
provided below the die opening to pre-
vent blanked parts or slugs from stack-
ing inside the entire width of the die
wall. The draft is commonly referred to
as angular clearance and is expressed
in degrees per side. A narrow cutting Fig. 9.23 The use of angular clearance
land is left below the cutting edge as
shown in Fig. 9.23. The cutting land, often called a die land, is 3.2 mm thick. On mate-
rials over 3.2 mm thick, the land is equal to the material. The main purpose of the die-
land width is to allow material for grinding the face of the during periodic sharpening.
Angular clearance is necessary to prevent or die block pressure caused by blank or
slug buildup especially when the punches or die block are fragile. Recommending
angular clearance varies from 0.25 to 2° per side, depending upon the material and
the shape of the workpiece. Soft materials and heavy-gauge materials require greater
angular clearance. Larger angular clearance may be necessary for small and fragile
punches.
Design of Sheet-Metal Blanking and Piercing Dies 609

Stripping The forces that cause the blank to grip inside the die walls also cause
the stock material to grip around the punch. The stock material will raise as the press
ram is raised unless some means of stripping the stock material from the punch is
provided. Figure 9.24 shows the two basic types of stripping device. Strippers are
discussed in detail in Section 9.11.
The amount of pressure required to strip the stock material from the punch varies
from 5 to 20 per cent of the cutting-force requirements. This is only a rough estimate,
as many variables affect stripping pressure. For example, some materials cling more
than others. Thicker materials requires more stripping force because more material
is in contact with the punch. Holes punched close to the strip edge do not require as
much stripping force because there is less backing and the metal can give. Punches
with polished sidewalls tend to strip easier than those with rough surfaces. More force
is also requires to strip punches that are close together.

Cutting forces The force required to penetrate the stock material with the punch is
the cutting force. If the die contains more than one punch that penetrates the stock
material simultaneously, the cutting force for that die is the sum of the forces for each
punch. Knowledge of cutting forces is important in order to prevent overloading the
press or failure to use it to capacity.

Fig. 9.24 Basic types of stripping devices: (a) Fixed (b) Spring-loaded stripper
610 Tool Design

The formula for determining cutting forces takes into account the thickness of the
stock material, the perimeter of the cut edge, and the shear strength of the stock
material. The shear of the stock material is the necessary to sever 1 sq mm of the
material by direct shearing action. Table 9.2 gives average shear strengths for vari-
ous materials.
The cutting force formula is
F = S pt
where F = cutting force
S = shear strength of stock material
p = perimeter or length of cutting edge
t = thickness of material
Table 9.2 Ultimate shearing strength of materials using flate-faced punch and die
Material Tons per N/mm2 Material Tons per N/mm2
sq. in. sq. in.

Aluminium, cast 6 92.52 Paper, using hollow die 1½ 23.13


Soft sheet 7½ 115.65 Using flat punch 4½ 65.53
Half-hard sheet 9½ 146.49 Bristol, flat punch 2½ 38.55
Hard sheet 12½ 192.75 Strawboard, flat punch 3½ 26.58
Asbestos millboard 2 30.84 Silver 15 231.29
Brass, cast 18 277.56 Steel casting 30 462.6
Soft sheet 15 231.29 Boiler plate 30 462.6
Half-hard sheet 17½ 269.85 Angle-iron 30 462.6
Hard sheet 20 308.4 Cold-drawn rod 29
Bronze, gunmetal 16 246.72 Drill rod, not tempered 40 616.8
1
Phosphor 20 308.4 Nickel about 3 __% 41 632.22
4
Copper, cast 12½ 192.75 Nickel about 5% 42½ 655.35
Rolled 14 215.88 Silicon 32½ 501.15
Cupronickel 20 308.4 Stainless 35 539.70
Duralumin, soft sheet 15 231.29 0.10 carbon (soft) 22½ 346.95
Treated 17½ 269.85 0.25 carbon (mild) 25 385.5
Treated and cold-rolled 20 308.1 0.50 carbon 35 539.70
Fiber, hard 12 185.04 0.75 carbon 40 616.8
Iron, cast 12½ 192.75 1.00 carbon 42½ 655.35
2% nickel 25 385.5 1.20 carbon, not tempered 47 724.7
Wrought 20 308.4 Tempered 95 1464.89
Lead 2 30.84 Hot*
Leather, chrome 3½ 54 Tin, cast 8 46.3
Oak 3½ 54 Rolled sheet 2½ 38.55
Rawhide 6½ 100.23 Sheet steel coated with tin 25 385.5
Monel metal, cast 30 462.6 Zinc, sand-cast 7 107.94
Rolled sheet 32½ 501.15 Die-cast 8 123.4
Nickel silver, half-hard sheet 16 246.72 Rolled sheet 9 138
Hard-rolled 10 154.20
*One-sixth to one-quarter strength of cold.
SOURCE: C W Hinman, “Die Engineering Layouts and Formula,’’ p. 450, McGraw-Hill, New York, 1943.
Design of Sheet-Metal Blanking and Piercing Dies 611

A 2-in. square hole is to be pierced in 1020 steel which is 0.062


Example 9.1 in. thick.
Solution: S = 25 tons/sq in. (0.25 carbon steel from Table 9.2)
p = 2 + 2 + 2 + 2 in. = 8 in.
t = 0.062 in.
Then,
F = 25 tons/sq in. × 0.062 in. = 12.2 tons
The answer to a cutting-force problem is generally expressed in tons because press
ratings are given in tons.
It sometimes becomes necessary to reduce cutting forces to prevent press overload-
ing. One method of reducing cutting forces is to step punch lengths as shown in
Fig. 9.25. Punches or groups of punches progressively become shorter by about one
stock-material thickness. A second method is to grind the face of the punch or die at
a small shear angle with the horizontal.This has the effect of the press and smoothes
out the cutting operation. The shear angle chosen should provide a change in punch
length of from 1to 1.25 times the stock thickness is called full shear. Cutting forces are
reduced by approximately 30 per cent when full shear is applied.

Fig. 9.25 Methods of reducing cutting forces: (a) Stepping punches (b) Single shear on punch
(c) Single shear on die (d) Type of double shear on dies (e) Types of double shear
on punches (f) Convex and concave shear

Varies types of shear are shown in Fig. 9.25. Double-angle shear is preferred over
single-angle shear because it does it does not set up lateral-forces components.
Double-shear angle on punches should be concave to prevent stretching the mate-
rial before it is cut. The shear angle may be applied either to the punch face or die
612 Tool Design

face, depending upon whether the operation is blanking or piercing because shear
will distort the work material. In other words, the shear angle for a blanking operation
will be on the die member, while in a piercing operation the shear angle will be on the
punch member.

Punch and die mountings A very simple set-up for a blanking or piercing operation
would be similar to that in Fig. 9.1. Here the punch is carried in the ram or slide, and
the die is bolted to the bolster. This set-up may be satisfactory for low-volume produc-
tion and wide tolerances, but where close tolerances are imposed, it is difficult to keep
the alignment between punch and die correct. Incorrect alignment not only causes
rejected workpiece but may cause the punch to strike the edge of the die.
To overcome this difficulty and also to economise on the use of expensive tool steel, it
is now almost universal practise to mount punches and dies on standardised die sets.
A die set is a pair of bases for mounting die
components that are accurately into the lower
base. Guide-pin bushings are mounted on the
upper base. An accurate slip fit is maintained
between the bushing and guide pin. Figure
9.26 shows a typical two-post, standardised
die set. In addition to aligning punch and die
members, using die sets holds set-up time
to a minimum because the die is installed as
a unit. Upon completion of a production run,
the die is removed and stored as a unit that
can be immediately placed into production
when the need arises. The die set can also
be separated and reassembled without dis- Fig. 9.26 A typical two-post stan-
turbing the relationship of the punch to die. dardised die set (Danly
This is important with regard to maintenance, Machine Specialties, Inc.)
as punch and die blocks can be sharpened
without removing them from the die set.
The various parts of a die set are shown in Fig. 9.27. The bottom base is called the die
shoe. The upper base is called the punch holder. Most punch holders in the smaller
die-set size are made with a shank that fits in the clamping hole in the lower end of
the punch-press ram. The shank diameter is determined by the press in which the tool
is to be used and must be specified on the tool drawing. The guide pins (also called
leader pins and guide posts) provide alignment alignment between the die shoe and
the punch holder. The guide bushings are mounted in the punch holder. The guide
bushings are mounted in the punch holder and slide over the guide pins. Bushing are
available in various lengths, of different materials, removable or press fit, and of plain
or ball-bearing type. The flange is a ledge that protrudes from the die shoe to provide
a means of clamping the die shoe to the bolster plate of the press. The available
surface for mounting punch and die components is called the die area. The die area
on the die shoe should be at least 6 mm larger all around the die block. If the bolster
opening is excessively lager, an oversize die holder should be used to bridge the
opening. A minimum clearance of 16 mm should be provide between punch and die
blocks and guide pins and bushings to allow clearance for the grinding wheel when
resharpening. When this is not possible, removable guide posts may be used for easy
to dies for maintenance purposes.
Design of Sheet-Metal Blanking and Piercing Dies 613

Fig. 9.27 Basic components of standard die set

Fig. 9.28 Standard die-set styles (Danly Machine Specialties, Inc.)

The shut height of the die set is the distance from the die shoe to the top of the punch
holder when the die is the shoe to the top of the punch holder when the die is in its
closed position. The shut height is established by the length of the guide pins, which
must be at least 12.5 mm shorter than the shut height in order to allow for reduced
shut height due to resharpening.
Many different sizes and styles of standard die sets are listed in manufacturs’ cata-
logues as shown in Fig. 9.28. The most frequently used style is the back-pin set, in
which the pins are at the back of the set, leaving a clear space for hand feeding blanks
for second operations. The good view of moving parts and freedom from obstructing
pins and bushings also aids the operator when feeding from left to right. If the load
is too heavy and the feed is from the front, the centre-pin type has advantages. In
this, the guide pins are in the alignment with the load along the trasverse centreline
614 Tool Design

of the set. When the load is very heavy and end feeding is required, the diagonal-pin
type can be used. In this one of the pins is placed at front of the pins along a diego-
nal line and leaving the ends clear for feeding. Generally the left pin is in front, but if
the designer wished to feed from the left, he can put the right pin forward. Deciding
which pin should be in front rests on which arrangement would give the best view of
the action of the stock stop and which would be safest and adaptable to standard
guards.
For round workpiece, especially for coining and shaving operations, round die sets
with pins at the back or along the centre can be used to advantage. This type has
round or approximately round punch holders and die shoes.
For roll-fed operations, especially for progressive dies with several stations, a four-pin
type should be used. This type provides maximum rigidity and accuracy of alignment
by having a guide pin at each corner of the die set. The front pins are inconvenient
for hand operation and may even be dangerous unless proper guards are fitted to the
press.
The material used in standard die sets is cast iron, cast steel, or rolled steel. Cast-
iron die sets are lower in cost because of the reduction in necessary machining but
are subject to cracking under heavy shock loads. They are primarily used for smaller
dies that do not require maximum strength. Cast steel has a greater toughness and a
greater resistance to shock loads. Rolled- steel die sets exhibit maximum toughness
and a greater resistance to shock loads and are generally used for larger and special
dies. When a large hole is to be machined through the die set for blank removal,
a steel die set should be used to increase strength, especially when the die set is
placed over a large hole in the bolster plate.
Manufacturers of die sets offer two grades of accuracy, precision and commercial.
The major difference between the grades is in the closeness of fit between bushings
and the guide pins. Tolerances between the bushings and guide pins are much closer
for precision-grade sets to assure extremely accurate alignment between punches
and corresponding holes in die blocks. The general rule for selection of grade is to
use commercial die sets for forming and drawing dies and precision die sets for cut-
ting operations. Precision die sets are also used when the die clearances are low or
when the die has fragile components that could break if precise alignment were not
maintained.
Typical dimensions of standard back-pin die sets are given in Table 9.3. In this series,
the guide pins are at the back of the set. There are three types in the series. The
regular, with an approximately square die space; the reverse, with dimension B sub-
stantially longer than dimension A; and the long narrow, in which dimension A is
longer and dimension B is narrow. This series can be used for a large number of
designs. Dimensions for other series/such as centre-pin sets, V-punch holder sets,
four-pin sets, round sets, and diagonal-pin sets can be obtained from manufactures’
catalogues.
When selecting and ordering die sets, the die area can be determined by use of full-
size templates obtainable from the manufacturer and marked to corresponding with
the catalogue number. When ordering, the following information must be specified:
1. Quantity and catalogue number
2. Type and length of bushing
Table 9.3 Nominal dimensions of flanged two posts die sets, metric equivalents in italic numerals for reference only (Courtesy: Die Set Engineer-
ing Handbook and Catalog, Lempco)
Design of Sheet-Metal Blanking and Piercing Dies 615

Contd...
Contd...
616 Tool Design

Contd...
Design of Sheet-Metal Blanking and Piercing Dies 617
Contd...
618 Tool Design

3. Overall length and type of guide pins, based on shut height and length of
stroke
4. Diameter of shank or no shank
5. Thickness combinations if other than those listed in the catalogue (special
thicknesses generally furnished at extra cost)

9.8 TYPES OF DIE CONSTRUCTION


Dies are classified by the type of operation performed and by type of construction
of the die. The classifications used throughout the die industry are too numerous to
mention in this text, and only a few of the common classifications will be discussed
in detail. Many of the classifications overlap, and it is often difficult to determine the
exact classification. For example, a cut-off die is often built into a progressive die to
chop up the scrap skeleton. By the same token, many progressive dies could be clas-
sified as drop-through dies because the blanked workpiece is dropped through the die
and press bed into a pan or other container.

Inverted dies On conventional dies, the punching element is generally mounted in


the punch hoder, and the die (containing the opening or cavity) is carried on the die
shoe attached to the bolster. Sometimes this arrangement is reversed and the punch
is mounted on the die shoe, as shown in Fig. 9.29. Note the knockout pin and the
combined pressure pad and stripper. This type of die is called an inverted die and its
advantage is that there is little possibility of thin blanks being bent, as they are in dies
of the type shown in Figs 9.30 and 9.1. In the upright die set, the blanks may become
crowed in the die cavity and thus jam. This is especially true if the press is working at
maximum capacity. In an inverted die, the blank is removed by means of a knockout
pin as it cut, instead of being forced through the die. The inverted die has the further
advantage that the cutting edge are kept clear of chips by the operation of the stripper
and ejector. These edges thus need less regrinding. The disadvantage of the inverted
die its relatively high cost.

Progressive dies Progressive dies perform two or more operations at different stages
every time the ram descends. The stock strip is advanced through a series of stations
that perform one or more distinct die operation on the workpiece. The strip must move
from the first through each succeeding station to produce a complete workpiece.
Thereafter, a complete workpiece is produced with each stroke of the ram.
The distance from one station to the next must be the same. The station-to-station
distance is also the same as the advance distance. The advance distance, called
advance for short, is the distance the strip moves in order to relocate (register) at
each successive station. The feed distance is the distance the stock strip is moved at
each stroke of the ram and may or may not be the same as the advance distance. The
reason the feed distance and advance distance may not be the same is that the strip
is often overfed a slight amount against a stop. Pilots then register the strip by pulling
it away from the stop. This prevents the stop from interfering in any way part registry.
The principle of the progressive die can be best explained with the aid of Fig. 9.30.
The stock strip is fed into the channel mechanically or by hand. The primary stop is
pushed in by hand, and the lead end is then fed into contact with it. The press is now
tripped to produce the pierced hole at station 1.
Design of Sheet-Metal Blanking and Piercing Dies 619

Fig. 9.29 Inverted die set

Fig. 9.30 Simple progressive die: (1) Stock strip (2) Die-stop activating pin (3) Primary die
stop (4) Blanking punch (5) Piercing punch (6) Punch plate (7) Stripper (8) Die
block (9) Die set (10) Automatic button die stop (11) Punch pilot
620 Tool Design

The primary stop is released, and the stock is advanced to station 2, where it con-
tacts the automatic button die stop. The stop pin in the button die stop is mounted in
a rubber grommet which allows the pin to float a slight amount. A second hit is made
(activation of the punch press), and the pilot on the blanking punch enters the pierced
hole and ensures exact alignment of the stock as the part is blanked. At the same
time the die-stop activation pin pushes the button die-stop pin below the edge of the
blank, and the strip is allowed to slide forward on the upward stroke of the ram. The
button die-stop pin returns to its normal position and catches the strip on the inside
wall of the blanked hole. A third hit is made and another complete part is produced.
Thereafter, a complete part is produced at each stroke of the press ram.
Progressive dies are often made with many stations. In some, blanks are cut at the
first station and the blank returned to the strip by means of spring plates. When estab-
lishing the sequence of operations for progressive dies, piercing operations must be
placed first. Advantage should be taken of any required holes in the workpiece for
piloting holes can be placed in the scrap part of the strip. Irregularly shaped punches
having frail projections that are hard to machine and likely to break after a few runs
should be avoided by punching out a portion of the blank at one station and finishing
it at another (see Fig. 9.31).

Fig. 9.31 Irregularly shaped progressive punches

Operations that require bending and forming must be done in the latter stations. If the
workpiece has been previously blanked, the blank may be returned to the strip and
carried by the strip to the next station for bending or forming. Case must be taken to
avoid having pierced holes too close to a bend.
The principle advantage of a progressive die is the number of operation that can be
achieved with one handling of the stock strip. The main disadvantage is that work-
piece may become “dished” as they are pushed through the die as they generally have
very little support. Thin stock of soft materials may cause trouble by bending or tearing
around piloting holes, especially in die sets having many stations where the friction
and inertia of the stock are considerable.

Compound die A compound die differs from a progressive die in that it performs
two or more cutting operations during one stroke of the press at one station only. In
order to do this, both the upper and lower member of the die set carry punching and
Design of Sheet-Metal Blanking and Piercing Dies 621

blanking elements which are directly opposed to each other. In other words, the pierc-
ing punches act in the opposite direction with respect to the blanking punch. A simple
compound die is shown in Fig. 9.32.

Fig. 9.32 Compound die

Figure 9.32 shows the compound die in a closed position. Note that the blanking
punch also serves as the piercing die. The sidewalls adjacent to the cutting edges of
the blanking-die opening are straight because the blank does not pass through the
die. The blank is return-ejected by the knockout mechanism that is actuated at the
return stroke of the press. A knockout bar is built into the ram of most pressure to
strike the knockout collar. Angular clearance must be provided in the piercing die to
allow slugs to drop through the die. The knockout plate is often used to support and
guide fragile punches.
Compound dies are slower in operation than progressive dies, but they have advan-
tages for certain jobs, especially where tolerances are close. (1) The cutting operation,
aided by the action of the knockout plate, ensure flatness of the blank (2), Pierced
holes in the workpiece (the blank) can be held to close tolerances with the edges. This
is very important when blanking out such parts as clock gears having a central hole,
(3) Large parts can be blanked in a smaller press if compound dies rather than pro-
gressive dies are used, (4) Progressive dies of necessity need long strip of material.
Sometimes scrap blanks are available, and these can be hand-fed to a compound die
under conditions where the saving in material will offset the cost of the labour.

Combination dies A die in which a cutting operation is combined with a noncutting


operation is referred to as a combination die. The cutting operations may include
blanking, piercing, trimming, and cutoff and are combined with noncutting operations
which may include bending, extruding, embossing, and forming. Figure 9.33 shows a
typical combination die that draws and blanks a shell.
622 Tool Design

Fig. 9.33 Combination die

Steel-rule dies Steel-rule dies were originally developed in the paper-box indus-
try. Basically they consisted of a thin, bevel-edged strips were referred to as steel
rules and were held in place by pressing them slightly into a slot cut in plywood was
mounted on the ram of a press and operated against a flat metal plate or hard wood
block. Paper-box blanks were produced like cookies cut with the common kitchen
cookie cutter. Later the principle of the steel-rule die was applied to the production of
printing and advertising materials, gaskets, auto-body lining panels, carpets, rubber
mats, insulation materials, and other metallic materials whose contour consisted of an
odd shape. Recent developments have included cutting aluminium and sheet-metal
parts.
Steel-rule dies are now classed as single-element and two-element dies. The single-
element die is the cookie-cutter type described above. It consists of a die board,
usually 16 to 20 mm plywood of five to seven ply generally of maple or gum with a
kerf sawed to the shape of the part. The rule is inserted into this kerf and extends 5 to
13 mm. It is recommended that the rule is inserted into at 90° ± 1/4°. Rule heights are
standardised at 23.34, 23.46 and 23.82 mm. They are usually hardened to RC52 to
55 after bending. This type of die is used mainly for banking softer materials such as
paper, cardboard, fibres, rubber, felt, leather, and similar materials. Stripper material
for these dies is neoprene rubber or cork sheet approximately 21 to 10 mm thick glued
inside the steel rule.
The two-element steel-rule die differs from the single-element die in that its cutting
action is similar to that of a conventional punch and die. The steel rule is formed,
heat-treated, and set into a kerf carefully sawed into a high-grade resin-impregnated
plywood is backed by a steel subplate or by the punch holder when the steel-rule die
is mounted in a die set. A male punch is mounted in the die shoe opposing the steel
rule, as shown in Fig. 9.35. The punch is constructed of die-board with 5 to 6 mm steel
plate shaped to the part outline mounted on top of it. Other punch elements and die
Design of Sheet-Metal Blanking and Piercing Dies 623

parts are added to the die for piercing holes and slots at the same time that blanking
is done. Pieces of die rubber or polyurethane are glued at appropriate locations to
serve as strippers.

Fig. 9.34 Two-element steel-rule die (J A Richards Company)

Steel-rule dies are considered as limited-production tools, although when correctly


designed and carefully constructed they may have a life of several hundred thousand
pieces when cutting steel. Their major advantage is that they can be fabricated at a
fraction of the cost of conventional dies. A steel-rule die designed for close-tolerance
624 Tool Design

blanking, for example, cost roughly about 20 per cent as much as a conventional die
made for mass production of similar work. Work can be located accurately to maintain
a tolerance of from ± 0.04 to ± 0.13 mm. Die design and fabrication are normally much
shorter than for a conventional die and require only moderately skilled workers.

Rubber-pad blanking This method of blanking employs a rubber pad in the punch
holder and a punch in the die shoe (see Fig. 9.35). It operates on the principle that
when rubber is compressed, it transmits the pressure in all directions. The rubber is
confined in a cavity in order to exert full force against the part. This type of blanking,
referred to as the Guerin process, is generally limited to the blanking of aluminium
alloy up to a maximum of 1 mm thick. The minimum hole diameter or width of cutout
is approximately 50 mm, and at least 38 mm trim is necessary for external cuts.

Fig. 9.35 Rubber-pad blanking by the Guerin process

Again referring to Fig. 9.35, a piece of sheet metal is placed on the steel die, and
the rubber pad is brought down into contact. Since the retainer is a rather close fit
over the platen, the rubber is completely enclosed and exerts full force against the
part. In the example shown, a hole is pierced and the excess is trimmed from around
the blank. A gripper plate must be placed around the punch to clamp the trim area
beyond the line of shear. This is to ensure that the metal does not form around the
cutting edge and prevent a clean shear. The Guerin process is best suited for use
on a hydraulic press because mechanical-press action is too rapid to permit proper
pressure build-up in the rubber. It is therefore considered a low-production or short-
run die, although production may be increased by the size of the rubber pad and the
number of punches and gripper plates. The main advantage is that the rubber pad
takes the place of many different die shapes.

Kirksite blanking dies Kirksite is a hard, tough zinc alloy often usec to replace tool
steel in short-run dies for blanking thin-gauge aluminium and soft-steel parts. It has a
high impact resistance and can be cast into the shape of the punch or die. Its surface
is dense and smooth and requires little machining before use. It can be quickly cast
into large shape and may be used on the day it was cast.
Figure 9.36 shows a simple Kirksite blanking die. The die block is made of Kirksite
and the punch of soft steel. In marking the die, the hole outline was cast or rough-
sawed slightly undersize. The punch was then forced through the die to “shear” it in
and ensure that the contour of the punch and die are alike.
Design of Sheet-Metal Blanking and Piercing Dies 625

The cutting edges of Kirksite dies tend to


be self-sharpening as it has a tendency
to flow plastically toward the cutting edge.
They also may have a tendency to spread
under certain conditions, and close toler-
ance cannot be held. However, worn dies
are quickly melted for reuse.

Fine blanking Fine blanking is a pro-


cess that has been developed in recent
years. The blanked workpieces produced
by this process are characterised by Fig. 9.36 A simple low-production blank-
smooth edges and a close wall thickness ing die made from Kirksite
between pierced holes and the peripheral
edge. Generally, no further machining operations are necessary to obtain blank or
hole edges comparable to machined edges. Sometimes abrasive tumbling or vibrat-
ing finishing are necessary to remove a slight burr on the blank.
Fine blanking is basically a method of blanking or piercing parts without die break.
This is done by controlling metal flow in the shearing area. A knife edge, or impinge-
ment ring, is pressed into the metal outside the cutting line and the metal outside this
line is restrained by application of great force. With the metal virtually unable to move
inward and a punch-to-die clearance between 8 mm and zero, a clean smooth-edge
cut can be made regardless of stock thickness.
A triple-action hydraulic press is required to get the holding pr gripping action, con-
trolled punch speed, and counterpressure on the part necessary for successful fine
blanking. It is also necessary to provide a controlled and slow entry of the punch
into the material to prevent the characteristic edge breakout encountered in nor-
mal blanking. The slug must be kept under tight pressure between the punch and
counterpunch.
The fine-blanking press cycle is shown in Fig. 9.37a, the bottom ram moves upward
to grip the stock under pressure and to impinge upon the metal. After impinge-
ment occurs, the bottom ram continues to a distance equal to the stock thickness
(Fig. 9.37b). Blanking and piercing are thus accomplished. Setting of the lower ram
can be controlled to stop within 25 mm when required. However, in some instances it
is desirable for the punch to enter the die. At Fig. 9.37c, the lower ram descends, and
an ejector removes the blank. Figure 9.37d shows the method of ejecting slugs when
holes are cut in the blank.
In short, tool construction is similar to that of a close-fitting compound blank-and-
pierce die. The press’s triple action permits coining and forming to approximately 30°
if required.
The strip, slug, and blanked part cannot be stripped or ejected before the tool opens
at least twice the stock thickness. This delay prevents the slugs from falling back into
the blanked part. Thin and small parts are blown out, whereas large thick ones are
removed mechanically.
The fine-blanking press must be of special design because there must be virtually no
play between the punch and die, and the hold-down force may be as high as 2.5 times
the blanking tonnage. The material for fine-blanked workpieces should have good
626 Tool Design

ductility to all proper metal flow. Fine-blanking dies are reported to cost approximately
50 per cent more than tools for conventional operations, but the cost is more than
offset by the elimination of a secondary trimming or finishing operation.

Fig. 9.37 Operational sequence in fine blanking: (a) Impingement (b) Blanking (c) Blank
ejection (d) Slug ejection

9.9 DIE-DESIGN FUNDAMENTALS


9.9.1 Blanking and Piercing Die Construction
The information on blanking and piercing dies thus far has been general in nature.
This section will deal with the specifics and will provide guidelines for die design and
construction.

Screws and dowels The components of dies are held together by socket-head cap
screws and are held in alignment by dowel pins. The head on the cap screw is almost
always recessed in a counterbored hole to eliminate projecting screw heads. Cap
screws used to secure die blocks are generally counterbored 3 mm deeper than the
cap screw head to allow additional material for die sharpening.
Design of Sheet-Metal Blanking and Piercing Dies 627

Socket-head cap screws and dowels are commercially available in a wide range of
sizes. Standard commercial dowels are finished to 5 mm larger than the nominal diam-
eter with a tolerance of ± 2.5 mm.
A minimum of one cap screw and two dowels are necessary to position and hold a
die commponent in place accurately. More cap screws may be used, but two, and only
two, dowels should be used for positioning. Most die designers try to use at least two
cap screws, but small components may allow only one because of space limitations.
The diameter of screws and dowels is also determined by the size of the component.
Generally, 10 mm screws are used on die components up to 150 mm square. Heavy
die components are usually secured with 12 to 16 mm-diam screws. Dowel diameter
should be the same be the same as the of the cap screws.
Dowels should be located diagonally across from each other and as far apart as pos-
sible to increase locational accuracy. All screws and dowels should be located from
1.5 to 2 times their diameter from the components edge. Whenever possible, screw
and dowel holes should be placed nearer the outer edge of the die block and as far
away as possible from the edges from the edges of blanking contour.
Dowel holes always extend through the die components so that dowels can be easily
removed. A hardened dowel pressed into a blind hole is almost impossible to remove
by dowel diameter, the dowel hole should be relieved as shown in Fig. 9.38a. This
practise is especially recommended when the dowel hole must be finished after heat
treatment, as it minimises lapping and fitting time.
The effective thread depth for screws be 1.5 times the screw diameter for general
applications and two times the screw diameter when the component is subjected to
shock loads. Whenever possible, threading hardened components should be avoided.
Die blocks, for example, should be clearance-drilled and counterbored to accept the
cap screws. The threads should be cut in the die shoe as shown in Fig. 9.38b.

Fig. 9.38 Application of screws and dowels: (a) Recessed dowel hole
(b) Cap screw in die block

Die-block design The design of the die block depends basically upon the workpiece
size and thickness, although the contours of the workpiece and type of die may be
influential at times. Die blocks for small workpiece are usually constructed from a solid
block of tool steel. The size of die blocks is generally based on the past experience of
the designer. However, the dimensions given in Table 9.4 will help the beginner.
The distance between the die opening and the outside edge of the die block should
be 1.25 times the thickness of the die block for smaller dies. This distance should be
628 Tool Design

Table 9.4 Die-Block Thickness for Mild Steel Strip


Strip thickness (mm) Die-block thickness (mm)

Up to 1.6 19 – 25
1.6 – 3.2 25 – 29
4.8 – 6.4 35 – 41
Over 6.4 41 – 51
Use thicker blocks for larger cutting perimeters. Thickness for soft nonferrous materials may be less; thickness for
alloys with higher shear strengths should be greater.

increased to 1.5 to 2 times the die thickness for large dies or when sharp corners are
present in the die-opening contour.
Solid die blocks that are symmetrical may be incorrectly assembled after repair.
This can be prevented by intentionally placing one dowel at a different distance from
its nearest screw hole, a practise often referred to as foolproofing the die block.
Substantial savings can be derived from using insert dies in the construction of die
block. Insert dies can be obtained from specialty companies as off-the-shelf items
it virtually any size needed. Insert dies are mounted in a soft die block as shown in
Fig. 9.39.
Die blocks made in two or more sections, known as sectional die blocks, are used to
conserve tool steel in the construction of large dies or when the complexity of the die
contour is such that it is easier to machine in sections. The die may also be sectioned
when the size of the die opening is too small to permit internal machining. An added
advantage of sectional die blocks is that only one component need be replaced in
case of die failure.
Sectional components may be screwed and doweled to a die holder or die shoe with
the sections butting against each other, as shown in Fig. 9.40. The sections should
be wide enough to resist tilting. Dies subjected to heavier cutting forces should be
constructed to resist lateral displacement of the die sections. The use of the sectional
components to the other components, as shown in Fig. 9.41, is advantageous for use
with moderate cutting forces. Extremely heavy cutting forces require that the sections
be nested in a pocket, as shown in Fig. 9.42.

Punch design The design of punches largely depends upon the area to be pierced
or blanked and the pressure required to penetrate the workpiece material. The area
to be pierced or blanked determines the method of mounting punches. For example,
punches for larger workpieces may be constructed from a solid block of tool steel and
bolted directly to the punch holder of the die set. On the other hand, small punches
may require a punch block for mounting the punch to the die holder. The punch must
also withstand the maximum blanking or piercing pressure. Small punches may
require punch support to prevent breakage.
The types of punches are too numerous to discuss in this text. Only basic types will be
considered: the plain punch, the pedestal punch, the perforator punch, and punches
mounted in punch plates.
Design of Sheet-Metal Blanking and Piercing Dies 629

Fig. 9.39 Commercial insert punches dies, guides, pilots, and methods of mounting
(Day Ton Progress Corp)

Fig. 9.40 Method of sectioning large die subjected cutting force

Fig. 9.41 The use of sectional components to tie other components


630 Tool Design

Fig. 9.42 Die sections nested in pocket to resist heavy cutting forces

Plain punches Plain punches are simply a block of hardened tool steel shaped to
conform to the cutting contour, as shown in Fig. 9.43. They are mounted directly onto
the punch holder of the die set or onto a flat punch plate if extra length is needed.
Screws and dowels secure plain punches in the same manner as solid die blocks.
The main advantage is ease and economy of construction.

Fig. 9.43 Plain punch constructed from solid block of tool steel

Plain punches must be large enough in area to allow for screws and dowels. Screw
and dowel holes should be located at least 1.5 to 2 times the screw or dowel diameter
from the cutting edge. Preferred screw diameters are not less than 10 mm, although
Design of Sheet-Metal Blanking and Piercing Dies 631

smaller screws may be used if pressures are not too high. Dowel diameters are the
same as those of screws.
The length and width of plain punches should be at least equal to the punch height to
ensure punch stability. When it is necessary to use heights greater than either punch
length or width, a different types of punch should be considered, especially if blanking
and piercing pressures are high or unbalanced.
Large plain punches may be sectioned in the same fashion as die blocks to conserve
material or to aid in the construction of complicated shapes.

Pedestal punches Pedestal punches, sometimes


called flanged punches, are constructed by machin-
ing the shape of the punch in a block of steel in a
manner that leaves a flange around the base of the
punch (see Fig. 9.44). A pedestal punch always
has a base area larger than its cutting-face area.
Its major advantage is its stability, caused by the
large base and the solid construction. Cutting
forces are dispersed through the large base, which
makes the pedestal design applicable for heavy
cutting loads.
The flange of the pedestal punch should be wide
Fig. 9.44 Pedestal-type punch
and thick enough to provide ample room for holes
used for mounting purposes. The flange should be blended into the cutting portion
of the punch by a large smooth fillet. The strength of pedestal punches with a small
cutting-face area can be increased by using the construction shown in Fig. 9.45.

Perforator-type punches Punches of this type may be considered insert punches


whose cutting-face diameter is approximately 25 mm or less. They may be fabricated
but as a general rule are purchased as standard commercial items. They are manu-
factured in a wide variety of shapes and sizes by companies who specialise in the
manufacture of insert punches, dies, guides, and pilots (see Fig. 9.36).

Fig. 9.45 Methods of strengthening pedestal punches


632 Tool Design

Figure 9.46 shows methods of mounting perforator (insert) punches. Where produc-
tion runs are small and there is a possibility of frequent changes, the type shown in
a may be used. A flat is machined on the body against which a setscrew bears. This
not only helps hold the punch but prevents rotation, which is important for noncircular
perforator punches. Back-up plates of hardened steel should be used in all punch
mountings similar to this in order to keep the punch heads from pressing into the soft
punch holder and thus working loose.

Fig. 9.46 Methods of mounting perforator-type punches

In b, the head fits securely into the recess in the punch plate, and not only is there
little likelihood of any looseness developing, but the punch can be accurately located
by dowelling the punch plate into position. Dimension a should equal 2b, and the
pressed portion should be made to a standard reamer size. Provision should be made
Design of Sheet-Metal Blanking and Piercing Dies 633

to prevent the rotation of noncircular punches. Various methods used by one manu-
facturer of standard perforators are shown in Fig. 9.39.
Small perforator punches may be mounted close to pedestal punches, as shown in c.
If the flange is hardened, as it generally is, the small perforator punch should not be
pressed directly into the flange. A soft metal plug, as shown, will help prevent crack-
ing of the flange.
Low-melting alloys can be used to mount punches, as shown in d. This rectangular
punches can be easily anchored in this manner. An alloy of bismuth, lead, tin, and
antimony with a pouring temperature of 177° C and the property of expanding after
solidifying is marketed under the name of Cerromatrix.
If frequent changing of punches is required in the same die set, as when a series
of runs for similar workpieces having different sized holes is being planned, the ball
retainer punch shown in e can be used to an advantage. The punch is firmly locked in
the punch plate by means of a spring-loaded ball pressing into a recess in the shank
of the punch. To remove the punch, the ball is pressed up against the spring by means
of a piece of drill rod inserted through the hold, and the punch is pulled out by hand.
To replace the punch, it is inserted into the retainer and pushed into place. The ball
automatically presses into its seat and locks the punch securely.
A method of mounting perforator punches to allow punch removal without removing
the punch plate is shown in f. A hole is drilled and tapped into the punch holder of
the die set directly above the punch. Set screws are then used to back the punch.
Punch replacement is then possible without removing stripper or punch plate. Since
the screws are hardened, any tendency for the punch to sink into the punch holder
is eliminated.

Punches mounted in punch plates Punch plates serve to hold, position, and in some
cases strengthen the punch. The perforator punch previously described is generally
mounted in punch plate. Perforators are easily mounted in punch plates because the
heads and shanks are round. Mounting holes are simply made with drills, reamers,
counterbores, or single-point boring tools. Rectangular or odd-shaped punches are
not so easily mounted in punch plates, however. Figure 9.47 shows various the meth-
ods of mounting punches of this shape.
The small punch in is mounted in a punch plate to increase its rigidity. The main pur-
pose of the punch plate is to locate and support the punch. No dowels are used, and
screws serve to prevent the punch from pulling from the punch plate. A slight interfer-
ence fit is used between punch and punch holder to ensure accurate location.
Figure 9.47b shows a step-head punch mounted in the punch plate. No screws or
dowels are used, as the step head prevents the punch from pulling from the punch
plate. An interference fit is used between punch shank and punch holder to ensure
adequate location. Figure 9.47c shows a bevel-head punch similar to a step-head
punch.
The method of holding the punch in d is sometimes used when the punch shape is
complicated. The open slot facilitates machining.
Rectangular or square punches are more difficult to mount in punch plates because
square comers must be machined into the punch plates. This inconvenience is often
overcome by machining clearance holes at he corners, as shown in Fig. 9.48a. The
634 Tool Design

Fig. 9.47 Methods of mounting punches in punch plates

time-consuming fitting of small protrusions on punches can be avoided in the same


manner as long as sufficient contact area remains around the punch perimeter (see
Fig. 9.48b).

Fig. 9.48 Using clearance holes to reduce fitting between punch and punch plate

Punch support In general, piercing punches should not be smaller in diameter than
the thickness stock they are to pierce. Where a small unguided punch must pierce
Design of Sheet-Metal Blanking and Piercing Dies 635

stock thicker than the punch diameter, the punch shank should be at least twice the
hole size in diameter and the cutting face should be ground to the hole size for a dis-
tance of about twice the stock thickness. Always avoid designing punches that would
have more than 100 mm of unguided length. A spacer block should be used between
the punch and punch holder (or punch plate) in order to shorten punches if more than
a 100 mm length is necessary.
Figure 9.49 shows the various methods of supporting slender punches. A quill is used
to strengthen slender punches, as shown in a. This construction has many advan-

Fig. 9.49 Methods of supporting slender punches

tages over turning or milling an integral punch of varied diameters. The punch unit
may be made of tool steel or purchased commercially, and in case of breakage or
wear it is easily replaced. It is assembled in the quill with a light press fit, and the quill
can be made of mild steel.
Slender punches can be supported and guided by sliding fits in the stripper plate, as
in b or c. For best results, the stripper plate should be hardened at the guide holes or
fitted with hardened bushings as in c. Guide bushings can be purchased as standard
items as shown in Fig. 9.39.
A quill-supported punch capable of piercing holes equal in accuracy and finish to
reamed holes is shown in d. The quill consists of two sleeves that intermesh when
three prongs of the upper sleeve slide in corresponding grooves in the lower one.
The use of the durable support sleeve gives guidance and strength to the punch and
eliminates problems inherent in using quills and stripper guide bushings, which leave
much of the punch unguided and unsupported. Extremely small holes with diameters
as little as one-half of the material thickness can be pierced with this system.
636 Tool Design

The use of more than one perforator in a single quill is shown in e. This practise per-
mits piercing small holes close together. Some means should be provided to keep the
quill from turning in the punch plate.

Punch shedders There are times during piercing or blanking operations when the
slug or blank tends to cling to the punch face and follow the punch face and follow the
punch out of the die opening. This is often referred to as slug pulling. During normal
operations, the slug clings to the die wall because of a slight amount of springback
in the blank or slug. When cutting conditions are correct, the amount of springback is
proportionate to the stock thickness and the area of the slug. Therefore, small holes
in thin material have very little springback and do not expand sufficiently to cling to the
die walls. Consequently, the slug may adhere to the punch face and be drawn out.
Another factor that contributes to slug pulling is the lubricant. Heavy lubricants make
slugs cling to the punch face. Slug pulling caused by heavy lubricants can be reduced
by changing to a lower-viscosity lubricant or to a water-soluble lubricant.
There are several common methods of preventing slug-pulling. They are generally
applied to then smaller sizes of perforator punches used in high-speed stamping
operations; however, they may be applied to larger punches when it is necessary to
break the oil bond between the slug and punch face. Figure 9.50 shows the various
methods. Spring-actuated shedder pins are located in the centre of the punch in a

Fig. 9.50 Various types of shedders used on punches (Day/Ton Progress Corp)
Design of Sheet-Metal Blanking and Piercing Dies 637

and b. In c, a rubber insert is placed in the face of the punch to obtain a similar effect.
When the shedder is used on larger punches for the sole purpose of breaking the oil
bond, its effectiveness is increased by off-centre mounting to increase the mechanical
advantage.
The use of air pressure to prevent slug pulling is shown at d. this method is more
complicated than spring-loaded shedder pins and also tends to blow lubricant away
from cutting areas. The air blast may be continuous on high-speed operations or timed
for slower operations.
A commercial perforator with a spring-loaded shedder pin is retractable for punch
sharpening and eliminates the need for disassembly as shown in e.
The use of a shedder pin is limited to punches of no less than 2.5 mm in diameter
or the equivalent. For punches below 2.5 mm in diameter double concave may be
ground on the face of the punch or a notch may be ground on the face of the punch.
This distorts the slug and helps to prevent it from clinging to the punch face.

9.10 PILOTS
The function of pilots is to position the stock strip accurately and bring it into proper
register for succeeding blanking and piercing operations. When the strip is fed by
hand, the strip stop allows a slight amount of overfeeding beyond the registry position.
A pilot then backs the strip into registry position in a direction away from the strip stop.
This prevents buckling the strip against the stop.
Mechanically fed strips are normally underfed and pulled forward in the same direction
as the feeding motion by the pilots because many mechanical feeding mechanisms
utilise a unidirectional locking device which prevents any backfeeding of the strip.

Pilot dimensions The pilot should fit the hole with a close sliding fit. Pilot size varies
from 0.05 to 0.08 mm smaller than the punch size for average work and from 0.012 to
0.025 mm for precision work. Pilot length is usually at least 6 mm longer than punches
in order to ensure that registry is complete before cutting begins. Pilot nose contours
are given in Fig. 9.51.

Fig. 9.51 Pilot nose contours: (a) Acorn type (b) Flattened-point types [Form ASTME,
“ASTME Die Design Handbook,” F W Wilson (ed.), McGraw-Hill, New York,
1965, by permission]

Pilot types Pilots are generally classified as direct or indirect pilots. Direct pilots,
sometimes called punch pilots, are mounted on the face of the punch, as shown in
Fig. 9.52. Some method of positive retention should be used to prevent the pilot from
dropping out of the punch. Press-fit pilots should be avoided for this reason.
638 Tool Design

Fig. 9.52 Methods of attaching pilots to punches

Indirect pilots are designed to enter previously pierced holes in the strip some dis-
tance away from the blanking punches. This practise provided more support under
the strip and helps prevent distortion (see Fig. 9.53). Spring-loaded pilots should be
used on materials over 1.6 mm thick, as shown in Fig. 9.54. This allows the pilot to
retract in case of a misfeed. Spring-loaded pilots are not necessary on thinner materi-
als because the pilot will pierce the strip rather than break in the event of a misfeed.

Fig. 9.53 The use of indirect pilots: (a) Standard pilots and punches, (b) Indirect pilot (Day/
Ton Progress Corp)
Design of Sheet-Metal Blanking and Piercing Dies 639

In this case, tapered slug-clearance holes through the die and lower shoe should be
provided.

Misfeed detectors Figure 9.55 shows a precision misfeed detector used in a man-
ner similar to that of a pilot. The detector senses out-of-register position of stock and
actuates switch to cut off the electric power to the press.

Fig. 9.54 Spring-loaded pilot Fig. 9.55 Misfeed detector (Day/


Ton Progress Corp)

9.11 STRIPPERS AND PRESSURE PADS


The purpose of a stripper is to remove the stock from the punch after a blanking
or piercing operation. Strippers are classed as fixed or spring-operated. Fixed strip-
pers are generally solidly attached to the die block or die shoe, while spring-operated
strippers travel up and down on the shank of a punch. A simple open-gap fixed strip-
per is shown in Fig. 9.1. The more common channel-type fixed stripper is shown in
Fig. 9.56.

Fig. 9.56 Channel-type stripper


640 Tool Design

Channel strippers This type is often the designer’s first choice because of simplic-
ity. The most common channel stripper consists of a rectangular plate mounted on
top of the die block. A channel or groove is milled, through which the strip is passed,
as shown in Fig. 9.56. The height of the channel should be 1.5 times the stock thick-
ness unless the strip must be lifted over a fixed pin stop. The width must be equal
to the strip width plus adequate clearance to allow for variations in the width of strip.
Stripper-opening clearance around the punch should be adequate to clear the punch
and should not be over one-half the thickness of the strip material.
The back edge of the channel may serve as a back gauge to correctly position the
strip in close-tolerance work. A stock pusher is used to hold the strip against the back
gauge, as shown in Fig. 9.57. The stock pusher shown is a commercial item consist-
ing of a hardened-steel roller mounted on a spring-loaded lever. It is adjustable to
three pressure with a standard spring.

Fig. 9.57 Roller stock pusher (Producto Machine Company)

When the back edge of the channel is used as a back gauge, it is diesirable to provide
a means of retarding wear. Small channel strippers may be constructed of tool steel
and heat-treated, but in most cases they are made from mild steel with inserts added
to increase wear resistance. Probably the simplest method is to press hardened
dowel pins along the back guide, as shown in Fig. 9.58a. Another mehod is to make a
separate back gauge from hardened tool steel as an integral par of the stripper (see
Fig. 9.58b). An advantage of this mehod is that the back gauge can be extended to
aid in aligning the stock strip for starting and feeding, as shown in Fig. 9.58c.
The thickness of the channel-stripper plate is determined to a large extent by the size
of the socket-head cap screws used to hold it in place. The stripper plate must be
thick enough to allow for the screw-head counterbores, which in most cases provides
adequate stripper strength. It addition to screws, dowels are used in order to ensure
accurate alignment on the die block.

Spring-operated strippers Spring-operated strippers, sometimes referred to as


pressure-pad strippers, employ spring to apply pressure to the stock strip. An advan-
tage of this type is that it tends to hold the strip flat during the press cycle. Figure 9.59
shows the principle of spring-operated strippers.
Design of Sheet-Metal Blanking and Piercing Dies 641

Fig. 9.58 The use of back gauges in channel-type strippers

Fig. 9.59 Spring-operated stripper


642 Tool Design

Spring strippers are commonly retained in suspension by socket-head stripper bolts,


a form of a shoulder screw available as standard items from die-specialty companies.
Figure 9.60 gives dimensions for standard socket-head stripper bolts.

Nominal D A H J T M R
screws size Shoulder Head Head Hexagon Key Head fillet Shoulder
or basic diameter diameter height socket engag- extension neck fillet
shoulder size ement diameter radius
diameter Max. Min. Max. Min. Max. Min. Nom. Min. Max. MIN
6.0 6.00 5.982 10.00 9.78 4.50 4.32 3 2.4 6.8 .25
8.0 8.00 7.978 13.00 12.73 5.50 5.32 4 3.3 9.2 .40
10.0 10.000 9.978 16.00 15.73 7.00 6.78 5 4.2 11.2 .40
12.0 12.000 11.973 18.00 17.73 9.00 8.78 6 4.9 14.2 .60
16.0 16.000 15.973 24.00 23.67 11.00 10.73 8 6.6 18.2 .60
20.0 20.000 19.967 30.00 29.67 14.00 13.73 10 8.8 22.4 .80

Nominal K F D1 G I N E
screws size or Shoulder Shoulder Basic Thread Thread Thread Thread Thread
Basic shoul- Neck Neck Thread Pitch Neck Neck Neck Fillet Length
der Diameter Diameter Width Diameter Diameter Width Radius
Min. Max. Max. Min. Min. Max. Max. Min.

6.0 5.42 2.5 M5 0.8 3.86 3.68 2.0 0.50 9.75 9.25
8.0 7.42 2.5 M6 1 4.58 4.40 2.5 0.53 11.25 10.75
10.0 9.42 2.5 M8 1.25 6.23 6.03 3.1 0.64 13.25 12.75
12.0 11.42 2.5 M10 1.5 7.89 7.66 3.7 0.77 16.40 15.60
16.0 15.42 2.5 M12 1.75 9.54 9.31 4.4 0.87 18.40 17.60
20.0 19.42 2.5 M16 2 13.20 12.96 5.0 1.14 22.40 21.60

Fig. 9.60 Metric dimensions of standard socket-head stripper bolts


(Courtesy: Holo-Krome Company)

Stripper springs are sometimes mounted over rods in place of stripper bolts (see
Fig. 9.61b). The rod should be long enough to support fully the spring’s ID in order to
minimise bending. The recommended way of doing this is to press a dowel of proper
length and diameter into one plate and to drill a clearance hole in the other plate to
allow the dowel to pass through.
Stripper springs may be contained in pockets as shown in Fig. 9.61c. The gap G of the
unsupported portion of the spring should not be greater than the spring diameter. For
optimum performance, the springs should be mounted in flat-bottomed pockets. Drill-
point bottoms in pockets do not provide flat support for the spring’s closed and ground
end and expose it to bending. Bending of this nature contributes to fatigue failure on
one side of the coils, and early breakage may be the result.
Design of Sheet-Metal Blanking and Piercing Dies 643

Do not use cap screws as rods as shown in Fig. 9.61d. This causes the spring to hang
up on one side during operation and places pressure on one or two coils instead of
distributing the pressure evenly over all coils.

Fig. 9.61 Methods of containing die springs


Holes for containing pockets should not be undersized. This causes binding of the
spring coils on the outside diameter and uneven pressure distribution. Pocket holes
should also be chamfered to prevent the individual coils from binding on the cor-
ners of the pocket. Pocketed diameters are generally 1.6 mm larger than the spring
diameter for springs ranging from 13 to 30 mm and 3 mm larger for larger diameters.
More specific information on pocket diameter and rod diameter may be obtained from
manufacturers’ catalogues.

Selection of stripper strings It is difficult to determine the pressure needed to effect


stripping. If the skeleton around the punch is frail and stretches easily, or even breaks
at several places, little pressure will be required to remove the scrap from the punch,
but if there is a substantial amount of material around several piercing punches, the
stripping pressure may be from 5 to 20% of the cutting pressure. The amount of pres-
sure needed to hold thin stock firmly while it is being cut must also be considered
when selecting springs. Most do not exceed 10% of the cutting force.
The cutting/perforating force, P, can be calculated using the following formula:
P=t×l×k N
where t = thickness of the material, mm
l = length/perimeter of shear, mm
k = shear strength of the material, N/mm2
Considering that the stripping force is 10% of the cutting force, the stripping force, Lst,
can be expressed as
644 Tool Design

Lst = t × l × k × 0.1 N
Once the stripping force is determined, the amount of space available for spring
mounting and number spring that can be fitted is then determined. These data are
essential for selection of proper die springs. The procedure for die spring selection
is based on several steps. The first step is decide on the level of production required
of the die-short run, constant production, etc. The next step is to determine com-
pressed spring length “H ” and operating travel “T ” from the die layout. Then select
the free length “C” as follows: Decide which load classification the spring should be
selected from - Light, Medium, Heavy, or Extra-Heavy Load. Then, choose the figure
nearest the compressed length “H ” required by the die design from the appropriate
charts. After selection of C, determine “X ” (initial compression) by using the following
formula: X = C-H-T. Then determine “R ” (total rate for all springs in newtons per mil-
limeter) by using the following formula:
Lst
R = ___
X
Divide “R ” by the number of springs to be used (if known) in order to get the rate per
spring. Then refer to the charts for springs having the desired rate. If the number of
springs is not known, divide “R ” by the rate of the spring you select for the correct
number of springs.
Select the proper stripper springs for a compound die. The
Example 9.2 workpiece material is 1020 steel and thickness is 3.25 mm, with
a blanking perimeter of 70 mm. Consider the free length of the
spring to be 51 mm.
Solution: Given: thickness, t = 3.25 mm, Perimeter of shear, l = 70 mm, free length,
C = 51 mm
The shear strength of 1020 steel is 385 N/mm2 (refer to Table 9.2, 0.25 Carbon)
Considering that the stripping force is 10% of the cutting force, the stripping force, Lst,
can be estimated as
Lst = 3.25 × 70 × 385 × 0.1 = 8758.75 N
Determine “X ” (initial compression) by using the formula: X = C-H-T
Considering the spring for heavy load the compressed length, H will be 41 mm (from
Table 9.5a), the operating travel, T may be 3.5 mm (slightly more than the thickness
3.25 mm)
Then, X = 51 – 41 – 3.5 = 7.5 mm
Determine “R ” (total rate for all springs in newtons per millimeter) by using the follow-
ing formula:
Lst 8758.75
R = ___ = _______ = 1167.83 N/mm
X 7.5
Considering that there are three springs on each side, therefore there will eight
springs in total.
1167.83
Therefore, R value for each spring will be = _______ = 145.98 N/mm
8
Then referring to the table 9.5d for spring having the desired rate, the following spring
may be selected:
This particular spring is selected because the rate of deflection of the spring is 156.6
N/mm, which close to the required deflection rate of 145.98 N/mm and the total deflec-
tion for the spring for long life is 10.2 mm, which is greater than the total required
deflection of 10 mm (7.5 + 3.5 mm).
Table 9.5a Chart for conversion of compressed length to free length for die springs

C Light Load Medium Load Heavy Load Extra Heavy Load


Free Length H-Compressed Length (mm) H-Compressed Length (mm) H-Compressed Length (mm) H-Compressed Length (mm)
(mm) Long Average Maximum Long Aveage Maximum Long Average Maximum Long Average Maximum
Life Life Deflection Life Life Deflection Life Life Deflection Life Life Deflection
25% 30% 40% 25% 30% 37.5% 20% 25% 30% 17% 20% 25%

25 19 18 15 19 18 16 20 19 18 21 20 19
32 24 22 19 24 22 20 26 24 22 27 26 24
38 29 27 23 29 27 24 30 29 27 32 30 29
44 33 31 26 33 31 28 35 33 31 37 35 33
51 38 36 31 38 36 32 41 38 36 42 41 38
64 48 45 38 48 45 40 51 48 45 53 51 48
76 57 53 46 57 53 47 61 57 53 63 61 57
89 67 62 53 67 62 56 71 67 62 74 71 67
102 76 71 61 76 71 64 82 76 71 85 82 76
114 86 80 68 86 80 71 91 86 80 95 91 86
127 95 89 76 95 89 79 102 95 89 105 102 95
140 105 98 84 105 98 87 112 105 98 116 112 105
152 114 106 91 114 106 95 122 114 106 126 122 114
178 133 125 107 133 125 111 142 133 125 148 142 133
203 152 142 122 152 142 127 162 152 142 168 162 152
229 – – – 172 160 143 – – – – – –
254 190 178 152 190 178 159 203 190 178 211 203 190
Design of Sheet-Metal Blanking and Piercing Dies 645

305 229 213 183 229 213 191 244 229 213 253 244 229
646 Tool Design

Table 9.5b Load deflection chart for light load springs (ISO colour – Green, High tensile
strength chrome silicon material, rectangular wire design)
Hole Rod Free RATE LOAD-DEFLECTION TABLE
Diam. Diam. Length Newtons Total Deflection Total Deflection Maximum Total Travel to
(mm) (mm) (mm) Reqd. to Recommended Recommended Operating Solid
A B C deflect for Long Life for Avg. Life Deflection
1 mm (25% of C) (30% of C) (40% of C)
Load Defl. Load Defl. Load Defl. Load Defl.
N mm N mm N mm N mm
25 11.0 69 6.3 83 7.5 110 10.0 157 14.3
32 8.8 70 8.0 84 9.6 112 12.8 162 18.3
38 7.4 70 9.5 84 11.4 112 15.2 164 22.0
44 6.3 69 11.0 83 13.2 111 17.6 164 26.0
10 5 51 5.4 68 12.8 82 15.3 109 20.4 158 29.0
64 4.5 72 16.0 86 19.2 115 25.6 170 38.0
76 3.7 71 19.0 85 22.8 113 30.4 172 46.0
305 0.9 65 76.3 79 91.5 105 122.0 152 178.0
25 19.1 119 6.3 143 7.5 191 10.0 258 13.6
32 16.5 132 8.0 158 9.6 211 12.8 303 18.3
38 13.7 130 9.5 156 11.4 208 15.2 303 22.0
44 11.6 127 11.0 153 13.2 203 17.6 304 26.0
12.5 6.3 51 10.2 130 12.8 155 15.3 207 20.4 307 30.0
64 8.2 131 16.0 157 19.2 209 25.6 312 38.0
76 6.2 118 19.0 142 22.8 190 30.4 276 44.0
89 5.3 119 22.3 143 26.7 190 35.6 276 52.0
305 1.5 112 76.3 135 91.5 179 122.0 257 175.0
25 31.5 197 6.3 236 7.5 315 10.0 422 13.3
32 23.5 188 8.0 225 9.6 300 12.8 385 16.4
38 21.0 200 9.5 240 11.4 319 15.2 443 21.0
44 17.5 193 11.0 231 13.2 308 17.6 432 25.0
16 8 51 16.3 208 12.8 249 15.3 332 20.4 477 29.0
64 12.6 202 16.0 242 19.2 323 25.6 459 37.0
76 10.3 196 19.0 236 22.8 314 30.4 456 44.0
89 9.3 207 22.3 248 26.7 330 35.6 497 53.0
102 8.2 210 25.5 252 30.6 336 40.8 506 61.0
305 2.6 200 76.3 240 91.5 320 122.0 484 184.0
25 56.0 350 6.3 420 7.5 560 10.0 703 12.6
32 42.7 342 8.0 410 9.6 547 12.8 678 15.9
38 33.8 321 9.5 385 11.4 514 15.2 640 18.9
44 28.4 312 11.0 375 13.2 499 17.6 632 22.0
20 10 51 24.9 317 12.8 380 15.3 507 20.4 641 26.0
64 19.3 308 16.0 370 19.2 493 25.6 619 32.0
76 16.1 306 19.0 367 22.8 490 30.4 633 39.0
89 13.5 300 22.3 360 26.7 480 35.6 610 45.0
102 11.9 304 25.5 364 30.6 486 40.8 622 53.0
114 10.5 299 28.5 359 34.2 479 45.6 624 59.0
127 9.3 295 31.8 354 38.1 472 50.8 609 66.0
140 8.5 297 35.0 357 42.0 476 56.0 618 73.0
152 7.9 299 38.0 359 45.6 479 60.8 634 80.0
305 3.8 288 76.3 346 91.5 462 122.0 601 159.0
Design of Sheet-Metal Blanking and Piercing Dies 647

Hole Rod Free RATE LOAD-DEFLECTION TABLE


Diam. Diam. Length Newtons Total Deflection Total Deflection Maximum Total Travel to
(mm) (mm) (mm) Reqd. to Recommended Recommended Operating Solid
A B C deflect for Long Life for Avg. Life Deflection
1 mm (25% of C) (30% of C) (40% of C)
Load Defl. Load Defl. Load Defl. Load Defl.
N mm N mm N mm N mm
25 12.5 25 107.2 670 6.3 804 7.5 1072 10.0 1315 12.3
32 80.9 647 8.0 777 9.6 1036 12.8 1265 15.6
38 64.8 616 9.5 739 11.4 985 15.2 1230 19.0
44 53.6 589 11.0 707 13.2 943 17.6 1190 22.0
51 46.4 592 12.8 710 15.3 947 20.4 1196 26.0
64 35.7 572 16.0 686 19.2 915 25.6 1148 32.0
6 29.4 559 19.0 671 22.8 894 30.4 1139 39.0
89 24.7 549 22.3 659 26.7 879 35.6 1115 45.0
102 21.2 540 25.5 648 30.6 865 40.8 1098 52.0
114 18.7 534 28.5 641 34.2 855 45.6 1085 58.0
127 16.8 534 31.8 641 38.1 854 50.8 1087 65.0
140 15.2 533 35.0 640 42.0 853 56.0 1100 72.0
152 14.0 532 38.0 639 45.6 852 60.8 1111 79.0
178 12.1 538 44.5 645 53.4 860 71.2 1119 93.0
203 10.5 533 50.8 640 60.9 853 81.2 1125 107.0
5 7.0 534 76.3 641 91.5 855 122.0 1129 161.0
38 101.4 963 9.5 1156 11.4 1541 15.2 1838 18.1
44 83.2 915 11.0 1098 13.2 1464 17.6 1765 21.0
51 71.3 909 12.8 1091 15.3 1454 20.4 1746 25.0
64 55.0 880 16.0 1056 19.2 1408 25.6 1700 31.0
76 46.1 875 19.0 1050 22.8 1400 30.4 1756 38.0
89 38.9 865 22.3 1038 26.7 1384 35.6 1739 45.0
102 33.6 857 25.5 1029 30.6 1372 40.8 1727 51.0
32 16 114 29.6 844 28.5 1012 34.2 1350 45.6 1718 58.0
127 26.3 834 31.8 1001 38.1 1335 50.8 1686 64.0
140 23.6 827 35.0 993 42.0 1324 56.0 1661 70.0
152 21.5 819 38.0 982 45.6 1310 60.8 1661 77.0
178 18.2 811 44.5 973 53.4 1297 71.2 1642 90.0
203 15.9 809 50.8 971 60.9 1294 81.2 1629 103.0
254 12.6 801 63.5 961 76.2 1281 101.6 1599 128.0
305 10.3 788 76.3 945 91.5 1261 122.0 1589 153.0
51 105.6 1346 12.8 1616 15.3 2154 20.4 2596 25.0
64 80.2 1283 16.0 1540 19.2 2053 25.6 2481 31.0
76 65.7 1248 19.0 1497 22.8 1996 30.4 2482 38.0
89 55.7 1239 22.3 1487 26.7 1983 35.6 2488 45.0
102 47.8 1219 25.5 1463 30.6 1951 40.8 2435 51.0
114 42.2 1203 28.5 1443 34.2 1925 45.6 2441 58.0
40 20 127 37.8 1201 31.8 1441 38.1 1922 50.8 2449 65.0
140 34.0 1189 35.0 1427 42.0 1903 56.0 2415 71.0
152 30.8 1171 38.0 1406 45.6 1874 60.8 2388 77.0
178 26.3 1169 44.5 1403 53.4 1870 71.2 2374 91.0
203 22.6 1147 50.8 1376 60.9 1834 81.2 2339 103.0
254 18.0 1145 63.5 1375 76.2 1833 101.6 2331 130.0
305 14.7 1122 76.3 1346 91.5 1795 122.0 2293 155.0
64 157.4 2519 16.0 3023 19.2 4031 25.6 4874 31.0
76 126.1 2396 19.0 2875 22.8 3833 30.4 4733 38.0
89 105.4 2346 22.3 2815 26.7 3753 35.6 4652 44.0
102 89.7 2287 25.5 2744 30.6 3658 40.8 4519 50.0
114 78.6 2241 28.5 2689 34.2 3586 45.6 4475 57.0
648 Tool Design

Hole Rod Free RATE LOAD-DEFLECTION TABLE


Diam. Diam. Length Newtons Total Deflection Total Deflection Maximum Total Travel to
(mm) (mm) (mm) Reqd. to Recommended Recommended Operating Solid
A B C deflect for Long Life for Avg. Life Deflection
1 mm (25% of C) (30% of C) (40% of C)
Load Defl. Load Defl. Load Defl. Load Defl.
N mm N mm N mm N mm
50 25 127 70.1 2224 31.8 2669 38.1 3559 50.8 4449 64.0
140 63.0 2207 35.0 2648 42.0 3531 56.0 4421 70.0
152 57.4 2183 38.0 2619 45.6 3493 60.8 4404 77.0
178 48.7 2167 44.5 2600 53.4 3466 71.2 4373 90.0
203 41.7 2115 50.8 2538 60.9 3385 81.2 4231 102.0
254 32.9 2091 63.5 2509 76.2 3345 101.6 4224 128.0
305 27.1 2070 76.3 2484 91.5 3312 122.0 4180 154.0
76 192.6 3660 19.0 4392 22.8 5856 30.4 6976 36.0
89 157.4 3503 22.3 4204 26.7 5605 35.6 6726 43.0
102 133.5 3403 25.5 4084 30.6 5445 40.8 6552 49.0
114 115.6 3294 28.5 3953 34.2 5271 45.6 6404 55.0
63 38 127 103.0 3270 31.8 3923 38.1 5231 50.8 6412 62.0
152 83.4 3168 38.0 3801 45.6 5068 60.8 6254 75.0
178 70.1 3117 44.5 3741 53.4 4988 71.2 6136 88.0
203 60.2 3057 50.8 3669 60.9 4892 81.2 6060 100.0
254 46.8 2969 63.5 3563 76.2 4751 101.6 5851 125.0
305 38.5 2938 76.3 3525 91.5 4701 122.0 5798 150.0

Table 9.5c Load deflection chart for medium load springs (ISO colour – Blue, High ten-
sile strength chrome silicon material, rectangular wire design)
Hole Rod Free RATE LOAD-DEFLECTION TABLE
Diam.(mm) Diam. Length Newtons Total Deflectio Total Deflection Maximum Total Travel
A (mm) (mm) Reqd. ton Recommended Recommended Operating to Solid
B C deflect for Long Life for Avg. Life Deflection
1 mm (25% of C) (30% of C) (37.5% of C)
Load Defl. Load Defl. Load Defl. Load Defl.
N mm N mm N mm N mm
25 16.3 102 6.3 122 7.5 153 9.4 188 11.6
32 14.0 112 8.0 135 9.6 168 12.0 224 15.9
38 11.7 111 9.5 134 11.4 167 14.3 228 19.4
44 9.8 108 11.0 129 13.2 162 16.5 224 23.0
10 5 51 8.6 110 12.8 132 15.3 165 19.1 224 26.0
64 6.8 108 16.0 130 19.2 163 24.0 221 33.0
76 5.7 108 19.0 130 22.8 162 28.5 226 40.0
305 1.3 101 76.3 122 91.5 152 114.4 205 154.0
25 28.9 181 6.3 217 7.5 271 9.4 366 12.7
32 22.6 181 8.0 217 9.6 271 12.0 365 16.1
38 19.1 181 9.5 218 11.4 272 14.3 381 19.9
44 16.1 177 11.0 213 13.2 266 16.5 372 23.0
12.5 6.3 51 14.0 179 12.8 214 15.3 268 19.1 378 27.0
64 11.0 177 16.0 212 19.2 265 24.0 366 33.0
76 8.8 166 19.0 200 22.8 250 28.5 343 39.0
89 7.5 166 22.3 200 26.7 250 33.4 344 46.0
305 2.2 166 76.3 199 91.5 248 114.4 349 161.0
25 55.7 348 6.3 418 7.5 522 9.4 629 11.3
32 40.3 322 8.0 387 9.6 483 12.0 547 13.6
38 35.2 334 9.5 401 11.4 502 14.3 621 17.6
44 30.5 335 11.0 402 13.2 503 16.5 646 21.0
16 8 51 27.0 344 12.8 413 15.3 516 19.1 670 25.0
Design of Sheet-Metal Blanking and Piercing Dies 649

Hole Rod Free RATE LOAD-DEFLECTION TABLE


Diam.(mm) Diam. Length Newtons Total Deflectio Total Deflection Maximum Total Travel
A (mm) (mm) Reqd. ton Recommended Recommended Operating to Solid
B C deflect for Long Life for Avg. Life Deflection
1 mm (25% of C) (30% of C) (37.5% of C)
Load Defl. Load Defl. Load Defl. Load Defl.
N mm N mm N mm N mm
64 21.0 336 16.0 404 19.2 504 24.0 651 31.0
76 17.7 336 19.0 403 22.8 504 28.5 682 38.0
89 15.2 339 22.3 407 26.7 509 33.4 690 45.0
102 13.3 339 25.5 407 30.6 509 38.3 685 52.0
305 4.1 315 76.3 378 91.5 473 114.4 630 153.0
25 90.2 564 6.3 676 7.5 846 9.4 925 10.3
32 68.1 545 8.0 654 9.6 818 12.0 880 12.9
38 54.8 521 9.5 625 11.4 781 14.3 855 15.6
44 45.2 497 11.0 596 13.2 746 16.5 810 18.0
51 38.9 496 12.8 595 15.3 744 19.1 801 21.0
64 30.3 485 16.0 582 19.2 727 24.0 789 26.0
76 24.7 469 19.0 563 22.8 704 28.5 768 31.0
20 10 89 21.4 475 22.3 570 26.7 713 33.4 790 37.0
102 18.6 473 25.5 568 30.6 710 38.3 795 43.0
115 16.3 468 28.8 562 34.5 702 43.1 780 48.0
127 14.5 462 31.8 554 38.1 692 47.6 777 53.0
139 13.1 456 34.8 548 41.7 685 52.1 774 59.0
152 12.1 459 38.0 551 45.6 689 57.0 772 64.0
305 6.1 462 76.3 554 91.5 693 114.4 802 132.0
25 12.5 25 166.2 1039 6.3 1246 7.5 1558 9.4 1649 9.9
32 124.7 998 8 1197 9.6 1496 12 1586 12.7
38 98.6 937 9.5 1124 11.4 1405 14.3 1505 15.3
44 83.2 915 11 1098 13.2 1373 16.5 1519 18.3
51 71.8 915 12.8 1099 15.3 1373 19.1 1528 21
64 55 880 16 1056 19.2 1320 24 1455 26
76 45.2 858 19 1030 22.8 1288 28.5 1445 32
89 37.8 842 22.3 1010 26.7 1263 33.4 1408 37
102 32.9 840 25.5 1007 30.6 1259 38.3 1405 43
115 29.2 841 28.8 1009 34.5 1261 43.1 1425 49
127 26.3 834 31.8 1001 38.1 1251 47.6 1422 54
139 23.6 822 34.8 986 41.7 1232 52.1 1419 60
152 21.7 825 38 990 45.6 1238 57 1417 65
178 18.4 818 44.5 982 53.4 1227 66.8 1399 76
203 15.9 809 50.8 971 60.9 1213 76.1 1386 87
305 10.5 801 76.3 961 91.5 1202 114.4 1357 130
38 166 1577 9.5 1893 11.4 2366 14.3 2532 15.2
44 136.4 1501 11 1801 13.2 2251 16.5 2447 17.9
51 116.1 1480 12.8 1777 15.3 2221 19.1 2397 21
64 87.7 1404 16 1685 19.2 2106 24 2238 26
76 70.9 1348 19 1617 22.8 2021 28.5 2178 31
89 59.9 1333 22.3 1599 26.7 1999 33.4 2163 36
32 16 102 51.8 1322 25.5 1586 30.6 1983 38.3 2153 42
115 46.1 1324 28.8 1589 34.5 1986 43.1 2185 47
127 41.5 1318 31.8 1581 38.1 1977 47.6 2214 53
139 37.5 1302 34.8 1563 41.7 1954 52.1 2202 59
152 34.2 1298 38 1557 45.6 1947 57 2191 64
178 29.1 1294 44.5 1552 53.4 1941 66.8 2175 75
203 25.2 1280 50.8 1536 60.9 1920 76.1 2163 86
254 20 1268 63.5 1521 76.2 1902 95.3 2147 107
650 Tool Design

Hole Rod Free RATE LOAD-DEFLECTION TABLE


Diam.(mm) Diam. Length Newtons Total Deflectio Total Deflection Maximum Total Travel
A (mm) (mm) Reqd. ton Recommended Recommended Operating to Solid
B C deflect for Long Life for Avg. Life Deflection
1 mm (25% of C) (30% of C) (37.5% of C)
Load Defl. Load Defl. Load Defl. Load Defl.
N mm N mm N mm N mm
305 16.6 1269 76.3 1522 91.5 1903 114.4 2151 130
51 170.6 2175 12.8 2610 15.3 3262 19.1 3390 19.9
64 128.7 2060 16 2471 19.2 3089 24 3210 25
76 105.3 2000 19 2400 22.8 3000 28.5 3224 31
89 87.7 1952 22.3 2343 26.7 2928 33.4 3129 36
102 76 1938 25.5 2326 30.6 2907 38.3 3143 41
115 66.4 1908 28.8 2290 34.5 2862 43.1 3081 46
40 20 127 59.5 1891 31.8 2269 38.1 2836 47.6 3102 52
139 53.6 1862 34.8 2235 41.7 2793 52.1 3057 57
152 48.9 1857 38 2228 45.6 2785 57 3072 63
178 41.5 1847 44.5 2216 53.4 2771 66.8 3054 74
203 36.1 1831 50.8 2197 60.9 2746 76.1 3038 84
254 28.9 1835 63.5 2202 76.2 2752 95.3 3083 107
305 23.8 1816 76.3 2179 91.5 2724 114.4 3033 128
64 211.9 3391 16 4069 19.2 5086 24 5305 25
76 167.4 3181 19 3817 22.8 4772 28.5 5024 30
89 139.8 3110 22.3 3731 26.7 4664 33.4 4932 35
102 121.9 3108 25.5 3730 30.6 4662 38.3 5042 41
115 107.2 3081 28.8 3698 34.5 4622 43.1 5041 47
127 94.6 3003 31.8 3603 38.1 4504 47.6 4914 52
50 25 139 85.5 2970 34.8 3564 41.7 4455 52.1 4935 58
152 77.9 2961 38 3554 45.6 4442 57 4945 63
178 66.4 2954 44.5 3544 53.4 4431 66.8 4966 75
203 57.4 2915 50.8 3498 60.9 4373 76.1 4905 85
229 51 2918 57.3 3501 68.7 4376 85.9 4926 97
254 45.7 2903 63.5 3483 76.2 4354 95.3 4943 108
305 37.7 2871 76.3 3445 91.5 4307 114.4 4913 130
76 304.7 5790 19 6948 22.8 8685 28.5 9398 31
89 250.4 5572 22.3 6687 26.7 8358 33.4 9143 37
102 211.9 5404 25.5 6484 30.6 8106 38.3 8965 42
115 185.6 5337 28.8 6405 34.5 8006 43.1 9032 49
127 164.1 5210 31.8 6252 38.1 7815 47.6 8908 54
63 38 152 132.9 5051 38 6061 45.6 7577 57 8729 66
178 111.7 4972 44.5 5967 53.4 7458 66.8 8608 77
203 96.3 4888 50.8 5866 60.9 7333 76.1 8520 89
229 85.5 4893 57.3 5871 68.7 7339 85.9 8647 101
254 76.9 4882 63.5 5858 76.2 7323 95.3 8735 114
305 63.4 4834 76.3 5801 91.5 7251 114.4 8742 138
Design of Sheet-Metal Blanking and Piercing Dies 651

Table 9.5d Load deflection chart for heavy load springs (ISO colour – Red, High tensile
strength chrome silicon material, rectangular wire design)
Hole Rod Free RATE LOAD-DEFLECTION TABLE
Diam. Diam. Length Newtons Total Deflection Total Deflection Maximum Total Travel
(mm) (mm) (mm) Reqd. Recommended Recommended Operating to Solid
A B C to deflect for for Deflection
1 mm Long Life Avg. Life (30% of C)
(20% of C) (25% of C)
Load Defl. Load Defl. Load Defl. Load Defl.
N mm N mm N mm N mm
25 22.1 110 5.0 138 6.3 165 7.5 205 9.3
32 17.5 112 6.4 140 8.0 168 9.6 209 11.9
38 16.3 124 7.6 155 9.5 186 11.4 272 16.7
44 14.0 123 8.8 154 11.0 185 13.2 276 19.8
10 5 51 11.9 121 10.2 152 12.8 182 15.3 268 22.0
64 9.6 123 12.8 154 16.0 185 19.2 276 29.0
76 7.6 116 15.2 144 19.0 173 22.8 248 33.0
305 1.9 114 61.0 143 76.3 171 91.5 250 134.0
25 41.3 207 5.0 258 6.3 310 7.5 457 11.1
32 32.9 211 6.4 263 8.0 316 9.6 472 14.3
38 27.1 206 7.6 258 9.5 309 11.4 477 17.6
44 23.3 205 8.8 256 11.0 307 13.2 485 21.0
12.5 6.3 51 20.0 204 10.2 255 12.8 305 15.3 476 24.0
64 15.2 195 12.8 244 16.0 293 19.2 441 29.0
76 13.5 205 15.2 256 19.0 307 22.8 507 37.0
89 10.9 193 17.8 242 22.3 290 26.7 455 42.0
305 3.1 189 61.0 236 76.3 284 91.5 451 145.0
25 75.5 377 5.0 472 6.3 566 7.5 712 9.4
32 60.9 390 6.4 488 8.0 585 9.6 776 12.7
38 48.7 370 7.6 463 9.5 555 11.4 744 15.3
44 43.3 381 8.8 476 11.0 571 13.2 825 19.1
16 8 51 35.9 366 10.2 458 12.8 549 15.3 758 21.0
64 28.9 370 12.8 462 16.0 555 19.2 787 27.0
76 24.5 373 15.2 466 19.0 559 22.8 830 34.0
89 20.8 371 17.8 464 22.3 556 26.7 833 40.0
102 18.2 372 20.4 464 25.5 557 30.6 833 46.0
305 5.8 356 61.0 445 76.3 534 91.5 806 138.0
25 239.9 1200 5.0 1500 6.3 1799 7.5 1879 7.8
32 180.4 1154 6.4 1443 8.0 1732 9.6 1811 10.1
38 144.0 1094 7.6 1368 9.5 1641 11.4 1770 12.3
44 120.0 1056 8.8 1320 11.0 1584 13.2 1743 14.5
20 10 51 101.2 1033 10.2 1291 12.8 1549 15.3 1651 16.3
64 77.1 986 12.8 1233 16.0 1480 19.2 1532 19.9
76 63.4 964 15.2 1205 19.0 1445 22.8 1545 24.0
89 53.9 960 17.8 1200 22.3 1440 26.7 1554 29.0
102 46.9 957 20.4 1197 25.5 1436 30.6 1561 33.0
114 41.5 946 22.8 1183 28.5 1420 34.2 1567 38.0
127 37.1 943 25.4 1179 31.8 1415 38.1 1571 42.0
140 33.8 946 28.0 1183 35.0 1420 42.0 1574 47.0
652 Tool Design

Hole Rod Free RATE LOAD-DEFLECTION TABLE


Diam. Diam. Length Newtons Total Deflection Total Deflection Maximum Total Travel
(mm) (mm) (mm) Reqd. Recommended Recommended Operating to Solid
A B C to deflect for for Deflection
1 mm Long Life Avg. Life (30% of C)
(20% of C) (25% of C)
Load Defl. Load Defl. Load Defl. Load Defl.
N mm N mm N mm N mm
152 30.8 937 30.4 1171 38.0 1406 45.6 1577 51.0
305 15.1 919 61.0 1148 76.3 1378 91.5 1560 103.0
25 12.5 25 376.5 1883 5.0 2353 6.3 2824 7.5 2782 7.4
32 285.5 1827 6.4 2284 8.0 2740 9.6 2883 10.1
38 222.4 1690 7.6 2113 9.5 2536 11.4 2676 12.1
44 190.9 1680 8.8 2100 11.0 2520 13.2 2875 15.1
51 156.6 1597 10.2 1996 12.8 2395 15.3 2583 16.5
64 121.0 1549 12.8 1936 16.0 2324 19.2 2540 21.0
76 99.8 1517 15.2 1897 19.0 2276 22.8 2596 26.0
89 84.1 1496 17.8 1870 22.3 2244 26.7 2558 30.0
102 73.2 1493 20.4 1867 25.5 2240 30.6 2597 35.0
114 65.0 1481 22.8 1852 28.5 2222 34.2 2626 40.0
127 58.0 1472 25.4 1840 31.8 2209 38.1 2601 45.0
140 52.5 1471 28.0 1839 35.0 2207 42.0 2625 50.0
152 48.2 1464 30.4 1830 38.0 2196 45.6 2645 55.0
178 41.2 1465 35.6 1831 44.5 2198 53.4 2676 65.0
203 35.9 1458 40.6 1822 50.8 2186 60.9 2666 74.0
305 24.2 1474 61.0 1843 76.3 2211 91.5 2839 117.0
38 390.5 2968 7.6 3710 9.5 4452 11.4 4542 11.6
44 318.7 2805 8.8 3506 11.0 4207 13.2 4424 13.9
51 269.7 2751 10.2 3439 12.8 4126 15.3 4342 16.1
64 204.9 2623 12.8 3278 16.0 3934 19.2 4119 20.0
76 165.8 2521 15.2 3151 19.0 3781 22.8 4076 25.0
89 140.3 2497 17.8 3121 22.3 3745 26.7 4119 29.0
102 121.0 2469 20.4 3086 25.5 3703 30.6 4087 34.0
32 16 114 106.3 2424 22.8 3030 28.5 3636 34.2 4063 38.0
127 95.8 2433 25.4 3042 31.8 3650 38.1 4151 43.0
140 86.3 2418 28.0 3022 35.0 3626 42.0 4126 48.0
152 78.6 2390 30.4 2988 38.0 3586 45.6 4104 52.0
178 66.7 2375 35.6 2969 44.5 3563 53.4 4072 61.0
203 57.8 2346 40.6 2933 50.8 3520 60.9 4048 70.0
254 46.2 2349 50.8 2936 63.5 3523 76.2 4114 89.0
305 38.2 2329 61.0 2911 76.3 3493 91.5 4075 107.0
51 364.3 3716 10.2 4644 12.8 5573 15.3 6037 16.6
64 268.0 3430 12.8 4287 16.0 5145 19.2 5483 230.0
76 218.9 3327 15.2 4159 19.0 4991 22.8 5568 25.0
89 183.9 3273 17.8 4092 22.3 4910 26.7 5627 30.0
102 158.7 3237 20.4 4046 25.5 4855 30.6 5570 35.0
114 140.8 3210 22.8 4013 28.5 4816 34.2 5717 41.0
40 20 127 125.2 3181 25.4 3976 31.8 4771 38.1 5662 45.0
140 112.6 3153 28.0 3941 35.0 4730 42.0 5618 50.0
152 103.3 3141 30.4 3926 38.0 4712 45.6 5711 55.0
178 88.1 3136 35.6 3920 44.5 4704 53.4 5755 65.0
203 76.7 3114 40.6 3893 50.8 4671 60.9 5779 75.0
254 60.6 3078 50.8 3848 63.5 4617 76.2 5742 95.0
305 50.3 3066 61.0 3833 76.3 4599 91.5 5782 115.0
64 423.8 5425 12.8 6781 16.0 8137 19.2 8467 20.0
76 338.0 5138 15.2 6422 19.0 7706 22.8 8273 25.0
Design of Sheet-Metal Blanking and Piercing Dies 653

Hole Rod Free RATE LOAD-DEFLECTION TABLE


Diam. Diam. Length Newtons Total Deflection Total Deflection Maximum Total Travel
(mm) (mm) (mm) Reqd. Recommended Recommended Operating to Solid
A B C to deflect for for Deflection
1 mm Long Life Avg. Life (30% of C)
(20% of C) (25% of C)
Load Defl. Load Defl. Load Defl. Load Defl.
N mm N mm N mm N mm
89 280.2 4988 17.8 6235 22.3 7482 26.7 8144 29.0
102 245.2 5002 20.4 6252 25.5 7503 30.6 8412 34.0
114 215.4 4911 22.8 6139 28.5 7367 34.2 8457 39.0
50 25 127 189.1 4804 25.4 6005 31.8 7206 38.1 8184 43.0
140 168.8 4727 28.0 5909 35.0 7091 42.0 7990 47.0
152 154.1 4685 30.4 5856 38.0 7028 45.6 8059 52.0
178 131.3 4676 35.6 5845 44.5 7014 53.4 8165 62.0
203 114.4 4643 40.6 5804 50.8 6965 60.9 8245 72.0
254 89.8 4564 50.8 5705 63.5 6846 76.2 8075 90.0
305 74.6 4551 61.0 5689 76.3 6826 91.5 8193 110.0

Table 9.5e Load deflection chart for extra-heavy load springs (ISO colour – Yellow, High
tensile strength chrome silicon material, rectangular wire design)
Hole Rod Free RATE LOAD-DEFLECTION TABLE
Diam. Diam. Length Newtons Total Deflection Total Deflection Maximum Total Travel
(mm) (mm) (mm) Reqd. Recommended Recommended Operating to Solid
A B C to deflect for for Deflection
1 mm Long Life Avg. Life (25% of C)
(17% of C) (20% of C)
Load Defl. Load Defl. Load Defl. Load Defl.
N mm N mm N mm N mm
25 32.7 139 4.3 164 5.0 205 6.3 241 7.3
32 25.6 139 5.4 164 6.4 205 8.0 238 9.3
38 21.2 137 6.5 161 7.6 201 9.5 243 11.5
44 17.7 132 7.5 156 8.8 195 11.0 230 13.0
10 5 51 15.4 134 8.7 157 10.2 196 12.8 230 15.0
64 12.3 133 10.9 157 12.8 196 16.0 233 19.1
76 10.2 131 12.9 154 15.2 193 19.0 236 23.0
305 2.5 127 51.9 150 61.0 187 76.3 231 94.0
25 58.7 249 4.3 293 5.0 367 6.3 468 8.0
32 44.1 240 5.4 282 6.4 353 8.0 431 9.8
38 36.3 234 6.5 276 7.6 344 9.5 433 12.0
44 30.6 229 7.5 270 8.8 337 11.0 434 14.2
12.5 6.3 51 27.0 234 8.7 275 10.2 344 12.8 460 17.0
64 21.7 236 10.9 278 12.8 347 16.0 486 22.0
76 17.7 229 12.9 269 15.2 336 19.0 469 26.0
89 15.1 228 15.1 268 17.8 335 22.3 465 31.0
305 4.2 219 51.9 257 61.0 322 76.3 450 107.0
25 127.3 541 4.3 637 5.0 796 6.3 1012 7.9
32 94.0 512 5.4 602 6.4 752 8.0 911 9.7
38 75.8 490 6.5 576 7.6 720 9.5 895 11.8
44 63.6 476 7.5 559 8.8 699 11.0 887 13.9
16 8 51 55.5 481 8.7 566 10.2 708 12.8 912 16.4
64 43.3 471 10.9 554 12.8 692 16.0 895 21.0
76 35.6 459 12.9 540 15.2 675 19.0 885 25.0
89 30.3 458 15.1 539 17.8 674 22.3 894 30.0
102 26.4 459 17.3 539 20.4 674 25.5 903 34.0
305 8.5 443 51.9 521 61.0 652 76.3 907 106.0
25 320.5 1362 4.3 1602 5.0 2003 6.3 2087 6.5
654 Tool Design

Hole Rod Free RATE LOAD-DEFLECTION TABLE


Diam. Diam. Length Newtons Total Deflection Total Deflection Maximum Total Travel
(mm) (mm) (mm) Reqd. Recommended Recommended Operating to Solid
A B C to deflect for for Deflection
1 mm Long Life Avg. Life (25% of C)
(17% of C) (20% of C)
Load Defl. Load Defl. Load Defl. Load Defl.
N mm N mm N mm N mm
32 239.9 1305 5.4 1536 6.4 1919 8.0 2050 8.5
38 194.4 1256 6.5 1477 7.6 1847 9.5 2099 10.8
44 161.8 1210 7.5 1424 8.8 1780 11.0 2073 12.8
51 139.6 1210 8.7 1424 10.2 1780 12.8 2105 15.1
64 108.8 1183 10.9 1392 12.8 1740 16.0 2101 19.3
20 10 76 89.7 1159 12.9 1363 15.2 1704 19.0 2137 24.0
89 75.7 1145 15.1 1347 17.8 1683 22.3 2106 28.0
102 65.3 1133 17.3 1333 20.4 1666 25.5 2080 32.0
114 57.4 1113 19.4 1310 22.8 1637 28.5 2063 36.0
127 51.7 1115 21.6 1312 25.4 1640 31.8 2088 40.0
140 46.6 1109 23.8 1304 28.0 1630 35.0 2071 44.0
152 42.6 1100 25.8 1294 30.4 1617 38.0 2059 48.0
305 21.0 1090 51.9 1282 61.0 1602 76.3 2085 100.0
25 12.5 32 353.8 1924 5.4 2264 6.4 2830 8.0 3153 8.9
38 280.2 1810 6.5 2130 7.6 2662 9.5 3080 11.0
44 231.2 1729 7.5 2034 8.8 2543 11.0 3049 13.2
51 197.9 1716 8.7 2019 10.2 2523 12.8 3027 15.3
64 153.8 1673 10.9 1968 12.8 2460 16.0 3044 19.8
76 125.0 1616 12.9 1901 15.2 2376 19.0 3008 24.0
89 105.4 1595 15.1 1877 17.8 2346 22.3 2983 28.0
102 91.1 1579 17.3 1858 20.4 2322 25.5 2966 33.0
114 80.9 1568 19.4 1845 22.8 2306 28.5 3029 37.0
127 72.2 1558 21.6 1833 25.4 2291 31.8 3010 42.0
140 65.6 1561 24.0 1837 28.0 2296 35.0 3053 47.0
152 60.2 1557 25.8 1831 30.4 2289 38.0 3093 51.0
178 51.3 1553 30.3 1827 35.6 2283 44.5 3102 61.0
203 44.7 1541 34.5 1813 40.6 2266 50.8 3112 70.0
305 29.6 1535 51.9 1805 61.0 2257 76.3 3131 106.0
38 488.6 3156 6.5 3713 7.6 4642 9.5 4860 10.0
44 404.6 3026 7.5 3560 8.8 4450 11.0 4930 12.2
51 345.0 2991 8.7 3519 10.2 4399 12.8 4979 14.5
64 266.2 2896 10.9 3407 12.8 4259 16.0 5068 19.0
76 215.4 2783 12.9 3274 15.2 4093 19.0 4987 23.0
89 182.1 2756 15.1 3242 17.8 4053 22.3 5030 28.0
102 155.7 2700 17.3 3176 20.4 3970 25.5 4892 31.0
32 16 114 135.7 2630 19.4 3095 22.8 3868 28.5 4764 35.0
127 121.9 2632 21.6 3096 25.4 3870 31.8 4849 40.0
140 111.0 2651 24.0 3119 28.0 3898 35.0 5018 45.0
152 100.9 2607 25.8 3067 30.4 3833 38.0 4929 49.0
178 85.6 2591 30.3 3049 35.6 3811 44.5 4886 57.0
203 74.6 2575 34.5 3029 40.6 3786 50.8 4941 66.0
254 59.5 2571 43.2 3025 50.8 3781 63.5 5017 84.0
305 49.6 2570 51.9 3023 61.0 3779 76.3 5068 102.0
51 558.7 4844 8.7 5698 10.2 7123 12.8 7815 14.0
64 422.1 4592 10.9 5402 12.8 6753 16.0 7675 18.2
76 338.0 4367 12.9 5138 15.2 6422 19.0 7551 22.0
89 280.2 4240 15.1 4988 17.8 6235 22.3 7276 26.0
102 243.4 4221 17.3 4966 20.4 6208 25.5 7604 31.0
114 213.7 4141 19.4 4871 22.8 6089 28.5 7560 35.0
Design of Sheet-Metal Blanking and Piercing Dies 655

Hole Rod Free RATE LOAD-DEFLECTION TABLE


Diam. Diam. Length Newtons Total Deflection Total Deflection Maximum Total Travel
(mm) (mm) (mm) Reqd. Recommended Recommended Operating to Solid
A B C to deflect for for Deflection
1 mm Long Life Avg. Life (25% of C)
(17% of C) (20% of C)
Load Defl. Load Defl. Load Defl. Load Defl.
N mm N mm N mm N mm
40 20 127 189.1 4084 21.6 4804 25.4 6005 31.8 7503 40.0
140 171.0 4066 24.0 4783 28.0 5979 35.0 7477 44.0
152 155.3 4014 25.8 4722 30.4 5903 38.0 7439 48.0
178 131.0 3975 30.3 4676 35.6 5845 44.5 7394 56.0
203 113.8 3928 34.5 4622 40.6 5777 50.8 7361 65.0
254 90.4 3902 43.2 4591 50.8 5738 63.5 7450 82.0
305 75.0 3886 51.9 4572 61.0 5715 76.3 7500 100.0
64 725.0 7888 10.9 9281 12.8 11601 16.0 12535 17.3
76 572.7 7399 12.9 8705 15.2 10881 19.0 12200 21.0
89 474.6 7181 15.1 8448 17.8 10560 22.3 11981 25.0
102 404.6 7015 17.3 8253 20.4 10316 25.5 11827 29.0
114 352.0 6822 19.4 8026 22.8 10032 28.5 11761 33.0
50 25 127 313.5 6768 21.6 7963 25.4 9953 31.8 11666 37.0
140 282.0 6722 23.8 7908 28.0 9885 35.0 11825 42.0
152 253.9 6562 25.8 7720 30.4 9650 38.0 11530 45.0
178 215.0 6503 30.3 7650 35.6 9563 44.5 11466 53.0
203 185.6 6406 34.5 7537 40.6 9421 50.8 11394 61.0
254 146.2 6314 43.2 7429 50.8 9286 63.5 11316 77.0
305 120.7 6257 51.9 7361 61.0 9201 76.3 11265 93.0

(Tables 9.5a to 9.5e, Courtesy: Danly IEM)

Hole Rod Free RATE LOAD-DEFLECTION TABLE


Diam. Diam. Length Newtons Total Deflection Total Deflection Maximum Total Travel
(mm) (mm) (mm) Reqd. Recommended Recommended Operating to Solid
A B C to deflect for Long Life for Avg. Life Deflection
1 mm (20% of C) (25% of C) (30% of C)

Load Defl. Load Defl. Load Defl. Load Defl.


N mm N mm N mm N mm
25 12.5 51 156.6 1597 10.2 1996 12.8 2395 15.3 2583 16.5

Rubber or urethane can be used as stripper springs. Figure 9.62 shows a standard
commercial stripper constructed of urethane, a tough elastic plastic which combines
the best features of rubber and plastic. It does not compress but flows and displaces
when a force is applied. The stripper shown changes shape and returns to its original
shape after the force is removed. This characteristic, along with the great strength
of urethane, provides a high stripping pressure. The urethane also flows around the
punch-point contour to strengthen the piercing point and dampen the shock.
656 Tool Design

Fig. 9.62 The use of urethane strippers (Porter Precision Products Company)

Urethene is not affected by oil or grease and will not mar or scratch the workpiece
material. The standard strippers shown are low in cost and are easily applied. Close
centre-to-centre punch mounting is possible. Table 9.6 gives dimensions of standard
urethane strippers. Urethane can also be purchased in pads, bars, or cylinders for the
construction of special stripers.

Guiding stock with spring-operated strippers Figure 9.63 shows various methods
for guiding the stock strip when using spring-operated strippers. Guide rails may be
mounted on the die block, as shown in a, b and c. At a, the stripper serves as pres-
sure pad and bears directly against the material. The guide rails do not contact the
stripper. When it is not practical to clamp the strip, guide rails may be used to prevent
the stripper fro clamping the work, as at b. Hooks are added to the guide rails at c in
order to contain the strip better.
Where space limitations are a factor, spool or button stock-strip guides may be used,
as shown in d. A series of at least three spool or button guides on each side of the
strip will satisfactorily guide the strip although they are not as efficient as guide rails.
Stock-guiding devices may be applied to the stripper plate as in the case of inverted
and compound dies (see Fig. 9.63e). Since the stripper plate is held in alignment only
by stripper bolts, it may be necessary to provide a means of aligning the stripper if
accurate strip positioning is necessary. This is usually accomplished by using two or
more alignment pins that work like guide posts on a die set (see Fig. 9.63f).

Stock stops and automatic stops As mentioned before, some method must be used
to locate the stock in the die set when hand feeding. Automatic-feeding mechanisms
usually do not require a stop within the die set because they can be adjusted to
advance the strip quite accurately.
The simplest form of stock stop is probably a dowel pin against which an edge of the
previously blanked opening is pushed after each stroke of the press (see Fig. 9.1).
Enough clearance is provided in the strip channel to allow the stock to be lifted above
the pin on the return stroke of the press and thus release the strip from the pin. It has
the disadvantage of demanding considerable skill on the part of the operator.
Design of Sheet-Metal Blanking and Piercing Dies 657

Table 9.6 Dimensions of standard urethane strippers (Courtesy: Dayton Progress


Corporation)
D d d1 t d2

10 18 23 6 1.6
13 23 26 6
16 28 31 6
20 33 36 7
25 40 43 7 3.0
32 50 55 7
38 60 65 8
40 60 65 8

D L Stripping Pressure (N) at Deflection of


3 mm 6 mm 9 mm

44 978 1401 —
54 734 1290 1512
10
64 703 1060 1268
74 670 1020 1220
44 1566 2647 —
54 1357 2180 2469
13
64 1081 1780 2158
74 811 1707 2139
44 2433 3513 —
54 1779 2958 3692
16
64 1526 2736 3202
74 1490 2650 3182
44 3002 4359 —
54 2580 3936 4581
20
64 2046 3424 4226
74 1939 3180 3980
44 4737 6605 —
54 3425 5515 6672
25
64 3291 5070 6205
74 3158 4781 5887
44 6383 9185 —
54 5693 8674 10008
32
64 4480 6961 8118
74 3469 6491 7570
44 8562 12521 —
40 54 6583 10497 12744
64 5804 9563 11453
658 Tool Design

Fig. 9.63 Methods of guiding stock with spring strippers

Figure 9.64a shows a trip stop. As the stock is fed forward, the pawl rises on the
ratchet principle, but when the operator pulls the stock back, the pawl drops and
locates the stock exactly against the vertical surface of the pawl.
When the workpiece is the same width as the stock and the feed is from one side
with no skeleton of scrap material passing out the other side, a shoulder stop is the
simplest effective type. This type is extensively used on progressive dies when the
last operation is a cutoff or trimming one. Figure 9.64b and c shows two types. The
one in b is wide as the stock and is fastened to the die. The one in c is fastened to the
stripper, leaving a clean space at the end of the set for the disposal of the workpiece.
The shoulder height should be such that the stop extends past the stock a distance
equal to at least twice the thickness of stock.
A starting stop, often called a primary stop, (Fig. 9.65) is used to position the stock
as it is initially fed into the die and is generally mounted on the stripper plate as
shown. In use, it is pushed inward by the operator with one hand and the end of
the stock strip is fed against it with the other. The first hit is completed and the stop
released. Thereafter, indexing of the strip is accomplished through an automatic stop.
Progressive dies may have more than one primary stop, depending upon the number
of stages preceding the automatic stop.
Design of Sheet-Metal Blanking and Piercing Dies 659

The primary die stop shown in Fig. 9.65 is a commercial item that may be purchased
at very low cost. It is made of annealed SAE 1070 steel that can be shaped and then
hardened.
A typical automatic die stop is shown in Fig. 9.66. In this, the pin-end finger is raised
by the trip screw as the punch descends and cuts the blank. On the return stroke, the
pin end of the finger drops, and the pin would drop into its former position if it were not
for the endwise action of the finger, which causes the pin to drop onto the top surface
of the stock instead of into the blank space. The mounting of the finger on the pivot is
loose enough to allow for this endwise movement. As the stock is fed forward, the pin
drops into the next blank, enabling the operator to locate the strip exactly. The use of
an automatic die stop of this type on the underside of stripper is shown in Fig. 9.67.

Fig. 9.64 Stock-strip stops: (a) Trip stop (b) and (c) Shoulder stops

Fig. 9.65 Primary (starting) die stop (Producto Machine Company)


660 Tool Design

Fig. 9.66 Automatic stop

Fig. 9.67 Automatic die stop on spring stripper


Design of Sheet-Metal Blanking and Piercing Dies 661

Automatic die stops should be purchased as standardised items. Typical examples


are shown in Figs 9.68 and 9.69. Figure 9.68 shows a Danly auto-gauge, where the
operator feeds the stock against the feed pin in the usual manner. The gauge pin is
raised by the action of the lever, but in this device, the gauge pin only rocks when
removed from the blank space. Thus, as the punch rises, the gauge pin is left on the
top surface of the stock and the stock is free to be fed forward until the gauge pin
drops into the next blanked hole.

Fig. 9.68 Standardised automatic stop

An unusual stop that requires no moving mechanisms (Fig. 9.70) is known as the
French stop and operates on the principle of cutting a shoulder in the edge of the
stock strip which acts as stop. A strip wider than necessary is inserted into the strip
channel until it contacts the shoulder built into the back gauge. The first hit of the
press performs the first station operation and at the same time punches from the side
of the strip a section of metal equal to the length of the advance. This operation leaves
a shoulder in the side of the strip. The strip is then advanced until the strip shoulder
contacts the shoulder on the back gauge on the return stroke of the ram.
The advantage of the French stop is its accuracy and speed of operation. It is espe-
cially well suited to light-gauge materials that are easily distorted when pushed or
pulled against a stop pin. Its main disadvantages are the extra cost of tooling and
extra stock scrap.
662 Tool Design

Fig. 9.69 Commercial automatic die stop (Producto Machine Company)

Fig. 9.70 Principle of the French stop

9.12 PRESSWORK MATERIALS


The chief materials used in presswork operations are mild steel, brass, copper, and
to a lesser extent mica, fibre, hard rubber, and celluloid.

Steel While a number of articles are made on punch presses from wire, rods, or
bars, the greatest production is from strip or sheet metal. Steel for this purpose is
rolled at the mill as a primary process. Eleven tempers (or hardness factors) are in
Design of Sheet-Metal Blanking and Piercing Dies 663

use. These are annealed, dead soft, soft, bright annealed, quarter hard, half hard,
three-quarters hard, and extra spring hard. Of these, the harder grades are used for
easy blanking and punching, and the dead soft for deep drawing.

Brass There are six tempers of brass, namely, annealed, quarter hard, half hard,
hard, extra hard, and spring hard. Cartridge brass (70 per cent copper and 30 per cent
zinc) is used extensively for deep-drawing operations.

9.13 STRIP LAYOUT


The die designer must consider the spacing of the blanks on the stock. There are two
main considerations: (i) the best location of blanks to save material and (ii) the best
location of blanks to serve good bending where bending is required. Figure 9.71a
shows a blank series where the percentage of stock scrapped would be too high.
Figure 9.71b shows a better arrangement for conservation of material. Figure 9.71c
shows how a slight change in workpiece design can result in maximum utilisation of
the stock strip.

Fig. 9.71 Location of blanks

When solving problems of strip layout, a cardboard or celluloid template cut to con-
form to the contour of the workpiece is often helpful. The outline of the workpiece is
traced from the template on a blank piece of paper. The template is then placed in
various positions adjacent to the workpiece outline until the most favorable relation-
ship is determined. The template must not be reversed or turned over unless a two-
pass die or a gang die is to be used.
The grain direction of the strip may be a consideration in the location of the workpiece,
especially if a bending operation is required. When sheet metal is rolled in the steel
mill, a fibre, or grain, is produced in the direction of rolling. To obtain maximum strength
from bent parts, the bends should be made at an angle of 90° to the grain direction.
The normal grain direction of a coil strip is parallel with the edge of the strip. When
workstrips are sheared from sheet stock, it is possible that the grain direction may
664 Tool Design

be 90° to the edge of the strip, depend-


ing upon how the strip is sheared from the
sheet.
The burr produced during the piercing
or blanking operation may determine the
position of workpieces in some instances.
On some workpieces, it may be desirable
to have the burr on one side. An example
would be to locate the burr on a hidden
surface and thus eliminate a deburring
operation. Remember that inverting the
blank position will also invert the burr side
with respect to the workpiece contour. Fig. 9.72 Strip layout by turning work-
Using workpiece templates may immedi- piece upside down and around
ately show the best strip layout with regard to material utilisation; however, in two or
three layouts the maximum utilisation may not be apparent. To find the per cent of
scrap or material utilisation, the following formulas may be used:
A–B
Per cent scrap = _____ × 100
A
B
Per cent utilisation = __ × 100
A
where A = area of uncut strip
B = total area of blank cut
In Fig. 9.72, the area of the uncut strip would be W × L., while the area of the blank
would be the area of the actual blank itself. In use, the per cent scrap or material utili-
sation is determined for each layout and compared for maximum material utilisation.
Irregular or odd-shaped blanks may pose the problem of lengthy calculations when
determining the area. The following formula may be used for comparison purposes
only:
random length of strip × width of strip
Total area used per blank = __________________________________________
number of blanks obtained in random length of strip
In using the above formula, the same random length is selected for each strip layout.
The area used per blank represents both the utilised and scrap material. The layout
with the lowest area per blank would be the best selection from a scrap standpoint.
The best strip layout is not always the one with the best utilisation of material. For
example, the layout with the best utilisation of material may require the construction
of a more complex die which would offset the savings obtained unless a large number
of parts is necessary.
It is always necessary to consider the spacing between blanks when laying out the
strip. Blanks located too close together or too close to the edge of the strip tend to
allow the metal to slip by the cutting edges of the punch and die. The web between
the blanks the forms the scrap skeleton must be strong enough to withstand feed-
ing forces, especially on a progressive die. A general rule of thumb for strip layout
is to make the web between the blanks and edges of the strip at least 1.5 times the
strip thickness; however, other factors may allow the web to be thinner, including the
Design of Sheet-Metal Blanking and Piercing Dies 665

thickenss of the strip, the hardness of the material, the length of scrap web, the shape
of the workpiece, and type of operation. Progressive and two-pass dies require wider
web thickness. Figures 9.73 and 9.74 give web thickness according to the above con-
siderations. Two-pass dies require webs that are approximately 50 per cent wider.

Fig. 9.73 Scrap-web allowances for webs with edges parallel (From D E Ostergaard, “Basic
Diemaking,” McGraw-Hill, New York, 1963, by permission)
666 Tool Design

Fig. 9.74 Scrap-web allowance between opposed curves or between curves adjacent to
straight lines (From D E Ostergaard, “Basic Diemaking,” McGraw-Hill, New
York, 1963, by permission)

9.14 SHORT-RUN TOOLING FOR PIERCING


Manufacturers offer as standard items adjustable punches and dies that can be quickly
set up for short production runs. In principle, adjustable punches and dies are similar
to a construction kit consisting of a universal die set and component parts. Their use
is probably the most inexpensive means of rapid tooling for simultaneous punching
of numerous holes in a variety of shapes and sizes. They are useful where multiple
holes are to be punched into materials on short runs and in some case in production
quantities as well. They save the time and money involved in punching or drilling holes
individually or in having an expensive permanent die made up. All component parts
are reusable. The speed with which adjustable punches and dies can be set up means
that they take only a fraction of the time required to make a permanent die, and the
skill of a toolmaker is not required in setting up.
Design of Sheet-Metal Blanking and Piercing Dies 667

Figure 9.75 shows the component parts which go into a typical adjustable punch and
die set. The setup is shown in Figure 9.76. In a, die-locating plugs are mounted in
die holders. Material holders with gauges inserted are also placed on the die set. All
locating plugs and stops have 6 mm points. A master template to establish location is
placed over die-locating plugs and material stop gauges in b. The die holders are then
fastened to the lower plate by special hold-down nuts.

Fig. 9.75 Component parts of a typical adjustable punch and die set (Di-acro, Division of
Houdaille Industries)

Fig. 9.76 The setup of an adjustable punch and die set (Di-acro,
Division of Houdaille Industries)
668 Tool Design

In c, punch-locating pins are secured in punch holders which are loosely mounted on
the top die shoe to permit exact positioning on die-locating plugs. Uniform clearance
is positioning on die-locating plugs, as shown in d. Punch holders are secured to the
top shoe by special hold-down nuts. All that remains in set-up is to replace punch-
locating pins and die-locating plugs with punches and dies, as shown in e. A stripper
plate of rubber strippers is mounted, and the die set is ready to be mounted on a press
for production in f.
Figure. 9.77 shows another type of flexible
tooling that can be changed in a matter of min-
utes. The punch and die shoe are permanently
fastened to the press, the punch shoe to the
ram, the die shoe to the bolster. Template-
locating posts are mounted in four corners of
each shoe. The punch retainers shave mag-
nets built into them and are held to the punch
shoe by magnetic attraction. Gravity holds the
die retainers down against the die shoe.
The desired workpiece hole pattern is jig-bored Fig. 9.77 The Whistler Magna-Die
into the upper and lower templates while they (S B Whistler & Sons, Inc)
are clamped together. This assures that the
two templates will be identical and that vertical alignment will be perfect. The function
of the templates is to locate the tooling within the confines of the die area. Figure 9.78
shows the set-up of the tooling and templates.

Fig. 9.78 Setup of the Whistler Magna-Die: (a) Die retainer complete with bushings being
inserted into a template that has been bored to a specific pattern (b) Snap rings
are placed around die bushings to hold retainers t template when inserting in die
set (c) Completely assembled punch and die templates, ready for insertion into die
set (d) Fastening template to die set. Die parts can be disassembled and regrouped
on different templates, eliminating a big investment in custom dies. (S B Whistler
& Sons, Inc.)
Design of Sheet-Metal Blanking and Piercing Dies 669

The templates are positioned over the locator posts mounted to each shoe of the die
set, and template clamp screws are fastened to each locator post to ensure safety
during operation.
The important feature of this system is that set-up is done internal to process time.
Change-over from one job to another simply consists of removing the templates from
the die set and quickly inserting the tooled templates for the new job to be run. Tooling
cost is amortised over all jobs put across this system.
The unit-tooling concept shown in Figs 9.79, 9.80, and 9.81 does not depend upon
the use of die set for alignment. This perforating and notching system consists of
mounting a selection of appropriate individual units by one of several mounting means
determined by the press equipment in use. The holder of each unit maintains perfect
alignment of punch and die at all times. Nothing is attached to the ram of the press,
which merely supplies the force to actuate the units. In some cases, however, con-
ventional die sets may be required to solve problem applications.
A pilot pin protruding from the base of each holder is exactly concentric with the
centreline of the punch and die contained in the unit, regardless of the size of the
hole perforated. This simplifies set-up since the positioning of each unit is controlled
by inserting the pilot pin of each unit into an accurately located hole in the mounting
template.
Figure 9.79 shows various components of the unit tooling concept. Figure 9.80 shows
a typical set-up on a press brake. A T-slotted press-brake bed rail is used for compo-
nent mounting. Any grouping of perforating and notching units having the same shout
height can be mounted together to perform a simple or complex operation in a single
stroke.

Fig. 9.79 Selected components for unit tooling (Pierce-all division,


Producto Machine Company)
670 Tool Design

Fig. 9.80 The use of unit tooling on a press brake (Pierce-all Division,
Producto Machine Company)

A flat-bed press is shown with unit-tooling components in Fig. 9.81. The area of perfo-
rating and notching is limited only by the press-bed area in this case. The units may be
mounted on a template or on a T-slotted bolster plate. When mounted on a template
the units can be unmounted at the completion of the run and templates placed in
storage while the unties are reassembled to another template to perform a different
operation.

Fig. 9.81 The use of unit tooling on a flat-bed press (Pierce-all Division,
Producto Machine Company)
Design of Sheet-Metal Blanking and Piercing Dies 671

Summary

The process of metal stamping is an integral and difficult part of manufacturing. The
requirement of accurate size blank and accurately pierced sheet metal enhances the
complexity of the process. This chapter deals with methods and devices used for
sheet-metal blanking and piercing and principles of designing sheet-metal blanking
and piercing dies.
• Different types of sheet metal die cutting operations are blanking, piercing,
lancing, cutting off, notching, shaving, trimming, etc.
• Presses are classified by (i) type of frame (ii) source of power, (iii) method of
actuation of sides, (iv) number of slides incorporated, and (v) intended use.
• The shearing action that occurs in blanking or piercing is quite similar to that of
chip formation ahead of a cutting tool.
• Clearance is defined as the intentional space between the punch cutting edge
and die cutting edge. Theoretically, clearance is necessary to allow the frac-
tures to meet when break occurs. The amount of clearance depends upon the
kind, thickness and hardness of the work material.
• Angular clearance is necessary to prevent backpressure caused by blank or slug
build-up especially when the punches or die block are fragile. Recommended
angular clearance varies from 0.25 to 2° per side, depending upon the material
and the shape of the workpiece.
• The force required to penetrate the stock material with the punch is the cutting
force. The cutting force F = Spt, where F = cutting force, S = shear strength
of stock material, p = perimeter or length of cutting edge, t = thickness of
material
• Different types of dies used for blanking and piercing are simple or conven-
tional dies, inverted dies, progressive dies, compound dies, combination dies,
steel-rule dies, Kerksite blanking dies, etc.
• The major components of a blanking and piercing die are: Screws and dowels,
punch, punch shedder, pilot, stripper and pressure pad, stock stop, etc.
• The designing parts to be blanked from strip material, economical stock utilsa-
tion is of high importance.

Questions
1. Be able to identify and understand the following basic die-cutting operations:
(a) Blanking, (b) Piercing, (c) Lancing, (d) Cutting off and parting, (e) Notching,
(f) Shaving and, (g) Trimming.
2. What is the advantage of the cast-frame construction of presses?
3. What is the advantage of welded-frame presses?
4. Why are gap-frame presses often manufactured with a separate frame to be
inclined at an angle?
5. What is the range in size of an OBI press?
6. What is the bolster plate of press?
7. What is the purpose of the knockout?
8. What is meant by a single, two, or four-point suspension of a press?
672 Tool Design

9. What are the characteristics of non-gear-drive presses?


10. What are the advantages of hydraulic-press drives?
11. How is the shut height of a mechanical press varied?
12. What is the major advantage of using an eccentric for driving a press?
13. What is meant by the action of a press?
14. What is meant by press tonnage?
15. How is the tonnage of a hydraulic press determined?
16. How is the tonnage of a mechanical press determined?
17. What is the shut height of a press?
18. What effect does the bolster plate have on the shut height?
19. What is the advantage of reel-type coil-unwinding equipment?
20. What is the purpose of the loop arm used on coil-winding equipment?
21. What is the primary function of a strip-straightening machine?
22. What is the basic operating principle of strip-straightening machines?
23. What is the advantage of a double-roll feed?
24. What determines the magnitude of the radius (sometimes referred to as a roll-
over) formed at the top edge of a pierced hole?
25. What causes the burnished band (often referred to as the cut band) that
appears on the walls of the pierced hole and slug?
26. In sheet-metal piercing and blanking what is meant by penetration?
27. How is the penetration of particular material expressed?
28. What is meant by die clearance?
29. How is meant by die clearance?
30. What is the result of excessive die clearance?
31. What is the result of insufficient die clearance?
32. What determines the correct amount of die clearance?
33. Why is die clearance increased for harder materials?
34. Is the die clearance placed on the punch or the die opening?
35. Is the die clearance placed on the punch or die opening for a blanking opera-
tion? Why?
36. Why is angular clearance necessary?
37. What determines the amount of angular clearance?
38. What is the amount of pressure required to strip the stock material from the
punches?
39. What factors contribute to the amount of force needed to strip material from
punches?
40. Why is the answer to cutting-force problems generally expressed in tons?
41. What methods are used to reduce cutting forces?
42. To what extent are cutting forces reduced by adding shear to punch?
43. When shear is added to punch, what determines the shear angle?
44. Why is double shear preferred over single shear?
45. What is the shut height of a die set?
46. Why should guide pins be 12 mm shorter than the shut height of the die set?
Design of Sheet-Metal Blanking and Piercing Dies 673

47. What materials are commonly used in the manufacture of die sets?
48. What is the major difference between precision and commercial die sets?
49. When are commercial-grade die sets sued? Precision grades?
50. What is the major advantage of the inverted die?
51. What is the difference between the feed distance and the advance distance of
a progressive die?
52. Is the feed distance greater or less than the advance distance?
53. What is the general rule for assigning forming and bending operations to sta-
tions in a progressive die set?
54. What are the major advantages of compound dies?
55. What is the major advantage of steel-rule dies?
56. What is the purpose of gripper plates used in rubber-pad blanking?
57. Why is a hydraulic press better suited for rubber-pad forming?
58. What is the advantage of using Kirksite as a tooling material in the construction
of short-run dies?
59. How is it possible to blank a part without die break by suing the fine-blanking
process?
60. What die clearances are used in the fine-blanking process?
61. What type of press is sued in the fine-blanking process?
62. What is the magnitude of the hold-down force used in fine blanking?
63. When are oversized dowels used?
64. What is the minimum number of screws and dowels necessary to accurately
position a die component and hold it in place?
65. What is the general rule for the size of screws and dowels for holding and
positioning die components?
66. Why should dowels be located diagonally across from each other and as far
apart as possible?
67. What is the general rule for locating screws and dowels with respect to the
component edge?
68. What is the effective thread depth for cap screws?
69. What is meant by foolproofing a die block?
70. What are insert dies?
71. Why are sectional dies used?
72. What is the general rule for plain-punch dimensions?
73. Why should perforator punches with small heads be backed up with plates of
hardened steel?
74. What is the general rule for piercing-punch diameter with respect to material
thickness?
75. What is meant by slug pulling?
76. How is slug pulling prevented on punches of less than 2.4 mm diameter?
77. Why is pilot length usually at least 6 mm longer than the punch length?
78. What is a direct pilot?
79. What is the advantage of indirect pilots?
674 Tool Design

80. Why should indirect pilots be spring-loaded when used on material over 1/16
1.6 mm thick?
81. What are the two general classifications of strippers?
82. Why are channel strippers often the designer’s first choice when selecting
strippers?
83. What is stock pusher, and how is it sued?
84. What is the advantage of spring-operated strippers?
85. Why should the use of standard twist drills to make pockets to contain springs
be avoided?
86. Why should spring pockets be bored oversize?
87. Why is it desirable initially to select stripper springs from the lowest load-rating
series?
88. What is the purpose of the primary stop (sometimes called starting stop)?
89. What is the unique feature and advantage of the French stop?
90. What tempers of steel strip are used for blanking? For deep drawing?
91. What are the major considerations when positioning blanks on the stock
strip?
92. Why is grain direction of the strip a consideration in locating the blank when a
bending operation is required?
93. What is the general rule of thumb for web thickness between blanks and the
edges of the strip?
94. What are the major advantages of short-run tooling for piercing operations?

Problems
1. Determine the proper die clearance for a 19 mm round punch and die that is to
be used on a universal ironworker the will pierce all types and thicknesses of
metal.
2. Determine the proper die clearance for the workpice shown in Fig. 9.82e.
3. Determine the proper die clearance for the wokrpiece shown in Fig. 9.82f.
4. Determine the force required to blank the workpiece shown in Fig. 9.83e.
5. Determine the force required to blank the workpiece shown in Fig. 9.84.
6. Determine by formula the force necessary to strip the stock material from the
punch when blanking the workpiece shown in Fig. 9.82d. Assume the cut
edges to be 0.5 mm apart and 0.5 mm form the strip edge.
7. Select the necessary springs for a spring-operated stripper when blanking the
workpieces shown in Fig. 9.83d. There is room for a 64 mm-long compressed
spring when the die is in the closed position.
Design of Sheet-Metal Blanking and Piercing Dies 675

Fig. 9.82 Blanked and pierced workpieces


676 Tool Design

Fig. 9.83 Blanked and pierced workpieces

Fig. 9.84 Strip layout for two-station progressive die, half-hard brass, 0.8 mm thick
Design of Sheet-Metal Blanking and Piercing Dies 677

Design Problems
1. Sketch a simple solid die block to blank the workpiece shown in Fig. 9.82c.
Show (a) die-block thickness, (b) distance between the die opening and the
outside edge of the die block, (c) location of holes for cap screws, (d) location
of dowel holes, and (e) type of tool steel and hardness after heat treatment.
2. Sketch a punch to blank the workpieces shown in Fig. 9.82c, 9.82b, 9.83c,
and 9.83e. Shown (a) type of tool steel and hardness after heat treatment,
(b) punch plate and method of punch to plate, (c) location of screw and dowel
holes (if they are used), and (d) length of punch.
3. Sketch a punch to blank the workpieces shown in Fig. 9.85a. Assume that the
punch will be mounted in a two-station progressive die with the two 3.2 mm
holes being punched in the first station. Incorporate direct pilots in the face of
the punch and show method of attaching.
4. Sketch a channel-type stripper to mount on the die block in Design Prob. 1.
Show (a) starting stop, (b) stock pusher, (c) method of preventing wear on back
gauge, and (d) automatic die stop.
5. Sketch a spring-operated stripper for one of the punches in Design Prob. 2.
Show in detail the method of containing springs.
6. Accurately sketch to scale the best strip layout for utilisation of material for the
workpieces shown in Figs 9.83d, and 9.84c. Show workpiece outlines in red.
Determine the percentage scrap for each layout.
7. Design a simple blanking die set to blank one of the workpieces shown in
Fig. 9.82, as assigned by your instructor. The drawing must be a complete
assembly with good drafting techniques in tool drawings as explained in
Chap. 1 Assume the available press has a hole diameter in the slide (ram) of
44 mm for the die-set shank and select a standardised die set. Show all dowels
and screws. Incorporate a satisfactory stock stop, stripper, and back gauge.
Give dimensions of shut height. Use standard size sheet of paper, make title
block, stock list and change block. Calculate the blanking pressure and die
clearance and show as a general note on the drawing. Be sure your lines are
dark enough to reproduce a print.
8. Design either a compound or progressive die to blank one of the workpieces
shown in Figs. 9.83 and 9.85, as assigned by your instructor. Follow the gen-
eral instructions given in Design Prob. 1.
678 Tool Design

Fig. 9.85 Blanked and pierced workpiece


Design of Sheet-Metal Blanking and Piercing Dies 679

References
Books
• ASM: “Metals Handbook,” vol. 4, 8th ed., “Forming,” Metals Park, Ohio, 1969.
• ASTME: “ASTME Die Design Handbook,” F W Wilson (ed.), McGraw-Hill, New York, 1955.
—: “Tool Engineers Handbook,” 2d ed., F W Wilson (ed.), McGraw-Hill, New York, 1959.
• Bredin, H W: “Tooling Methods and ideas,” The Industrial Press, New York, 1967.
• Eary, D F and E A Reed: “Techniques of Pressworking Sheet Metal,” Prentice-Hill, Engelewood
Cliffs, N.J. 1958.
• Ostergaard, E: “Basic Diemaking,” McGraw-Hill, New York, 1963.
—: “Advanced Diamaking,” McGraw-Hill, New York, 1967.
• Paquin, J R: “Die Design Fundamentals,” The Industrial Press, New York, 1962.
• Schubert, P B: “Die Methods,” The Industrial Press, New York, book 1, 1966; book 2, 1967.
• SME: “Fundamentals of Tool Design,” Prentice-Hall, Englewood Cliffs, N. J., 1962.
Periodicals
• American Machinist, New York
• Machine and Tool Bluebook, Wheaton, Ill
• Machinery New York
• Manufacturing, Engineering, and Management, Dearborn, Mich
• Modern Machine shop, Cincinnati, Ohio
• Tooling and Production, Cleveland, Ohio
Catalogues
• Danly Machine Corp., Chicago.
• Dayton Profress Corp., Dayton, Ohio.
• Di-Acro Division, Houdaille Industries, Lake City, Minn.
• Dickerson Company, Springfield, Mass.
• Do All Company, Des Plaines, Ill.
• F. J. Littell Machine Co., Chicago.
• Forter Precision Products, Cincinnati, Ohio.
• Producto Machine Company, Bridgeport, Conn.
• J. A. Richards Company, Kalamazoo, Mich.
• Superior Die Set Corp., Oak Creek, Wis.
• The verson Allsteel Press Co., Chicago.
• S. B. Whistler and Sons, Inc., Buffalo, NY.
DESIGN OF SHEET–METAL
10 BENDING, FORMING AND
DRAWING DIES

10.1 INTRODUCTION
Chapter 9 dealt with using dies to cut sheet metal. This is only half the story when
considering sheet-metal press working operations. Most products of sheet-metal
presswork are not flat blanks that can be produced by die-cutting operations alone.
They generally have a third dimension obtained by a shaping operation which deform
the metal to various degrees. The shaping operation may be performed in the same
die as the cutting operation, or it may be done separately.
Shaping operation are generally divided into three groups, based upon how the
parent metal flows or deform during the shaping process, namely, bending, forming,
and drawing. Bending, the simplest of the three, is defined as shaping metal around
a straight axis which extends completely across the material. The result is a plane
surface at an angle to the original plane of the workpiece material. Metal flow is
uniform along the bend axis, with the inner surface of the bend in compression and
the outer surface in tension.
Forming is similar to bending, except that the form or bend is along a curved axis
instead of a straight one. A part that is formed generally takes the shape of the punch
or die. Metal flow is not as uniform as in bending because it may be localised to some
extent, depending upon the shape of the workpiece. The metal flow in forming dies
is not severe enough to affect the thickness or total area of the workpiece material to
any great extent. Bending along a large radius in a straight line may also be referred
to as a forming operation.
Drawing is the most complicated of the three sheet-metal shaping processes. A draw-
ing operation begins with a flat blank which is transformed into a cup or shell. The
parent metal is subjected to severe plastic deformation. Shell forms produced may be
cylindrical or rectangular with straight or tapered sides. Figure 10.1 gives examples of
the three sheet-metal shaping processes. Each is discussed in detail in the following
sections.

Fig. 10.1 Basic operations for shaping sheet metal: (a) Bending (b) Forming (c) Drawing

10.2 BENDING DIES


Bending operations are generally planned in order to orient the workpiece so that
bends are made across the grain produced in the rolling mills. This allows sharper
Design of Sheet–Metal Bending, Forming and Drawing Dies 681

bends without danger of cracking. When bends are made in two directions, it is desir-
able to orient the workpiece so that no bends are parallel with the grain direction. The
minimum bend radii are limited by several factors, including the angle and length of
bend, material properties and direction of bend in relation to grain. As a rule of thumb,
the minimum bend radii for most annealed (soft) metals is equal to the thickness of
the metal. However, material thicknesses require proportionately larger bending radii
than softer materials. Minimum bend radii can be obtained for specific materials from
handbooks.

Bending methods The two bending methods used extensively in presses are wip-
ing dies and V bending. V bending is accomplished by using a V block for a die and
a wedge-shaped punch to force the metal into the die (see Fig. 10.2). The desirable
width of the opening in the V is ordinarily at least eight times the material thickness, up
to about 16 mm plate. Die opening up to 10 or 12 times are used for forming heavier
thicknesses of plate. The included angle of the V bend can be changed by varying the
distance the punch forces the sheet metal into the V die. When the punch does not
“bottom out,” the processes are referred to as air bending. In other words, the material
being formed contacts only at the point of the male die and two edges of the V-die
opening. Air-bend dies are made with sharper angles that the angle to be formed. For
making right-angle bends in mild steel, both punch and die are generally made with
an included angle of 89o. Included angles of from 30 to 70o may have to be used for
some heat-treated and springy materials.

Fig. 10.2 Methods of bending sheet metal: (a) V bending (b) Wiping die

When 90° bends are made with air-bend dies,


the metal will form to a larger radius when a
wider-opening V is provided in the lower die.
This radius is approximately equal to one-
eight the width of the die opening, even though
the die opening is wider than eight times the
metal thickness. As long as the punch has a
nose radius slightly smaller than the metal
thickness, the radius formed on the part will
depend entirely on the width of the opening
in the die.
The advantage of air bend dies is that a wide Fig. 10.3 Press brake used for V bend-
range of angles can be formed with one set ing (Verson Allsteel Press
of dies and each set is capable of bending Company)
682 Tool Design

different types and thickness of materials. Air-bend dies also permit bends to be made
with the least possible pressure, since all the ram pressure is used in bending the
metal rather than squeezing it. The dies do not have to be used in matched sets. For
example, one male die can be used with more than one size V-die opening. Air-bend
dies are primarily used on press brakes, as shown in Fig. 10.3. Figure 10.4 shows
typical press-brake bending and forming dies.

Fig. 10.4 Typical press-brake bending and forming dies [From ASTME “ASTME die Design
Handbook,” FW Wilson (ed.,) McGraw-Hill, New York, 1965, by permission]

When punch is allowed to bottom out in a V-bend die, the tool is referred to as bot-
toming die. In this case, the stock material is squeezed between the male and female
die surface at the bottom of the press stroke. Bottoming dies are used when a higher
degree of accuracy or shaper inside corner are required. Bottoming is also a means of
Design of Sheet–Metal Bending, Forming and Drawing Dies 683

reducing springback. Bottoming dies require from 3 to 10 times as much pressure as


air-bend dies, depending upon the design and thickness of material. Maximum pres-
sure is required to produce the minimum inside radius. Bottoming dies must be used
in matched pairs, according to the thickness of the stock to be formed and the radius
needed. Care must be exercised to prevent overloading the press when bottoming
dies are used. Presses used for this type of V bending are usually equipped with a
tonnage indicator to prevent overload.
A wiping die is shown in Fig. 10.2b. Here the workpiece is clamped to a die block by a
spring-loaded or fluid-cylinder pressure pad, and the punch wipes the extended mate-
rial over one edge of the die. A bend radius is provided on the edge of the die block.
A radius or chamfer is often provided on the leading edge of the punch to prevent the
wiping action from being too severe.

Springback Elastic stresses remaining in the bend area after bending pressure
is released will cause a slight decrease in the bend angle. Metal movement of this
type is known as springback, and the magnitude of movement will vary according to
material type, thickness, and hardness. A larger bend radius will also cause greater
springback.
Overbending is the simplest way of combating springback problems, especially in
V-die air bends. The workpiece is bent through a greater angle than required, and the
workpiece spring back to the required angle. Springback for low-carbon and soft non-
ferrous material is from 0 to 2°. For 0.40 to 0.50 carbon steel and half-hard materials
springback may vary from 3 to 5°. Springback may be as high as 10 to 15° in harder
materials. These figure should be used only as an approximation because of other
variable that influence springback. The only practical way to determine the necessary
amount of overhead is trial and error.
Another method of preventing springback is by coining (squeezing) the metal slightly
at the corner in order relieve elastic stresses. This is sometimes referred to as corner
setting. The metal in the immediate corner is made to flow plastically and set up com-
pression strains that overcome elastic stresses. Figure 10.5 shows various methods
of modifying the punch nose for corner-setting operations. Corner setting verges on
bottoming out, and pressure is built up rapidly; for this reason large contact areas
should be avoided. Only a small amount of squeezing is necessary. Generally, com-
pression of metal from two to five per cent of the metal thickness will be satisfactory.
Springback can be prevented in wiping dies by “ironing” the material. To iron the bend
effectively the distance between the punch and die must be slightly less than the
metal thickness. When severe ironing is required, a backup heel may be necessary
to support the punch, as shown in Fig. 10.6a. Ironing in the corner will be effectively
accomplished only by using a slight shoulder on the punch that contacts the work
material near the end of the stroke (see Fig. 10.6b). Severe ironing may cause metal
to tear and will reduce the metal thickness, depending upon the amount of interfer-
ence between workpiece and punch. The die may also be undercut to permit over-
bending when springback is not too severe, as shown in Fig. 10.6c.

Bend allowance In bending operations, the material near the bend radius is under
compression while the material near the outside of the bend is under tension as shown
in Fig. 10.7. A neutral plane exists between the area under tension and area under
compression. If the material is uniform in section and its elastic limit is not exceeded,
684 Tool Design

Fig. 10.5 Methods of corner setting to prevent springback

Fig. 10.6 Methods of preventing springback in wiping dies

the neutral plane will coincide with the centreline of the material. However, when the
bending forces cause the elastic limit of the material to be exceeded, the neutral plane
Design of Sheet–Metal Bending, Forming and Drawing Dies 685

moves toward the inner surface at a distance of one-third to one-half the thickness of
the material. This is because the material under compression resists bulging much
more than the material under tension resists stretching, and therefore, the greatest
amount of plastic flow takes place on the tension side of the bend. The length of the
neutral plane does not change as a result of bending.

Fig. 10.7 Bend terms for general angle

Fig. 10.8 Typical channel bend

When a blank or sheet is to be bent, it is necessary to consider the effect of stretch-


ing the metal at the outside of the bend. Since there is no stretch in the neutral plane,
the length of the formed part along the neutral plane will be the correct length. The
curved neutral plane of the bend area is called the bend allowance as shown in
Fig. 9.7. Several methods and formulas are used for determining the bend allowance.
The “ASTME Die Design Handbook” gives
A
B = ____ 2p (IR + Kt)
360
where B = bend allowance (arc length of neutral axis), mm
A = bend angle, deg
IR = inside radius of bend, mm
t = metal thickness, mm
K = constant of neutral axis location
K is equal to 0.33 when IR is less than 2t and 0.50 when IR is more than 2t.
686 Tool Design

Example 10.1 Calculate the blank length of the part shown in Fig. 10.8.
Solution: Divide the blank into segments and number as shown in Fig. 10.8.
Determine the length of the segments
L1 = 50 – (5 + 3) = 42 mm
90
L2 = ____ (2p) (5 + 0.33 × 3) = 9.41 mm
360
Since IR is less than 2T, K = 0.33,
L3 = 100 – (5 + 3 + 5 + 3) = 84 mm
L1 = L5
L2 = L4
The blank length = L1 + L2 + L3 + L4 + L5 = 42 + 9.41 + 84 + 9.41 + 42 = 186.82 mm.
Note that the formula takes into account the stock-material thickness, the size of the
bend angle, and the size of the bend radius. These factors, along with the ductility of
the material, have the major influence on bend severity.
U dies and channel dies This type of tooling for bending, which may be considered
a cross between V-bending and wiping dies, is so named because the workpieces
produced in them bear a resemblance to the letter U or a channel (see Fig. 10.9).
U dies are generally equipped with a pressure pad, as shown, which helps prevent
the metal from bowing away from the flat face of the punch. This will not completely
eliminate springback, and it may be necessary to bottom out against the pressure
pad to prevent the legs of the channel from springing apart on the return stroke. This
additional pressure will increase the press load be several tons but produces a part
with the bottom of the channel only slightly hollow at its corner.

Fig. 10.9 The principle of the U die (Channel die)

Materials that are springy may be corner-set as


shown in Fig. 10.5d, to aid in preventing springback.
It is also helpful to release the sides of the punch 1
or 2°.
Figure 10.10 shows a method of controlling spring-
back on hardened material when corner setting would
cause a fracture at the corner. A concave radius
is machined in the punch, and a matching convex
radius is machined on the pressure pad, allowing a Fig. 10.10 Preventing spring-
slight amount of bending at the corners. back in U dies
Design of Sheet–Metal Bending, Forming and Drawing Dies 687

Bending pressures The amount of pressure required to bend the workpiece mate-
rial depends upon the thickness of the stock, die opening, length of bend, and the
amount of coining, bottoming, or ironing used. Air dies require the least pressure of
all bending methods. To estimate the pressure required for bending with air die, the
following formula, derived from the beam formula may be used:
KLSt2
F = _____
W
where = bending force required, N
K = 1.33 (die-opening factor based on a die opening of eight times
metal thickness; use 1.20 when die opening reaches 16 times metal
thickness)
L = length of bent part, mm
S = ultimate tensile strength, N/mm2
t = metal thickness, mm
W = width between contact points on die (die opening on V die; see
Fig. 10.11a)
For U bending and channel bending, the constant K = 0.67. Thus, the formula would
be
0.67LSt 2
F = ________
W
W is determined as shown in Fig. 10.11b.
For a die performing a single wiping bend, the constant K = 0.333. The formula would
be
0.333LSt 2
F = _________
W
W is determined as shown in Fig. 10.11b.

Fig. 10.11 Determining W for bending-pressure formula: (a) Air bending (b) Channel
bending and wiping dies

Calculate the bending force for 45° bend in 24ST3 aluminium,


Example 10.2 1.6 mm thick and 1.22 m long, with a die opening eight times
the metal thickness. The bend is to be made by air bending
methods.
Solution: The expression for calculating the bending force, F is given as
688 Tool Design

KLSt 2
F = _____
W
As the die opening is 8 times the metal thickness, the die opening factor, K = 1.33
The ultimate tensile strength for the material, S = 448 N/mm2
The die opening width, W = 8 × t = 8 × 1.6 = 12.8 mm
Therefore, the bending force

1.33 × 1220 × 448 × 1.62


F = _____________________ = 145.385 KN
12.8
Calculate the bending force for the channel bend shown in
Example 10.3 Fig. 10.8. The bend length is 1 m, and die radius is 10 mm.

Solution: The expression for calculating the bending force, F is given as


KLSt 2
F = _____
W
The die opening factor, K for channel dies = 0.67
Let us consider the ultimate tensile strength for the material, S = 380 N/mm2
t = 3 mm (Ref. Fig. 10.8)
The die opening width W = R1 + R2 + C
= 10 + 10 + 3 = 33 mm (Ref. Fig. 10.11)
Therefore, the bending force
0.67 × 1000 × 380 × 32
F = ___________________ = 179.02 KN
12.8

10.3 FORMING DIES


It is sometimes difficult to distinguish between a bending and forming die. As stated
at the beginning of the chapter, a forming operation is generally along a curved axis
rather than a straight axis. Often the shape of the punch and die is reproduced in the
metal. There is not enough metal flow in a forming operation to cause excessive thin-
ning or tearing of the metal.
Forming operations may be simple or extremely complicated. It is difficult to develop
general formulas for forming pressures because of the many workpiece shapes which
fall under the forming classification. The shape of the part may also dictate the need
for pressure pads in the design to hold the metal in place during the forming operation.
This is especially true if small-radius corners are called for in the workpiece design.
In addition to shaping the metal to the desired workpiece contour, forming operations
may strengthen the workpiece and add rigidity, remove sharp edges and improve the
appearance of the workpiece in general. Forming operations may be done to facilitate
assembly later in the production line. The particular type of forming operation classi-
fies forming dies in groups. Each group has its own name, which describes the shape
of the formed section. The more common ones are as follows:
Design of Sheet–Metal Bending, Forming and Drawing Dies 689

1. Solid form dies 2. Pad-type form dies 3. Curling dies 4. Embossing dies
5. Coining dies 6. Bulging dies 7. Assembly dies

Solid form dies Forming dies of this type consist of a male punch and female die
shaped to the contour of the workpiece. Allowance for clearance equal to the thick-
ness of the workpiece is provided between the male and female die. This type of
tooling is generally used for forming operations in progressive dies. Solid form dies
are mounted at one or more of the later stations in the progressive die to shape the
workpiece after blanking. The form punches and dies are mounted in die sets in the
same manner as punches and die blocks at the blanking stations. Since solid form
blocks are very close to bottoming out in the closed position, careful consideration
must be given to the possibility of overload in case of misfeed, double feed, or varia-
tion in material thickness. Some designers call for the form die block to be made in
two pieces to help control cracking in case of overload. The sections are able to
spread and give slightly in a direction away from the punch. Die sections should be
nested or keyed together, as explained under sectional dies in Chapter 9. Screws and
dowels used in holding and locating form die blocks should be as large as possible
and still be in proportion to the size of the die block because of heavy lateral forces
that can be expected in form dies.
In piercing and blanking dies, the workpiece material has someplace to go because
there is an opening through which the material can pass. This is not the case in form
dies, and therefore side pressure builds up. The use of screws and dowels to resist
all lateral forces must be restricted to light work. Heavier work dictates that the die be
constructed in a manner that will provide adequate resistance to side thrust. This is
generally accomplished by nesting or keying the blocks.
Screws and dowels are preferably located about 1.5 times their diameter from the
edge of the form die block. They should be spaced properly and located so they will
not mar the workpiece.
When the shape of the workpiece is such that the work material must move over a
form edge, a radius should be provided on the form edge to prevent digging into the
material and galling. The radius should not be less than twice the material thickness,
and the edge should have a high polish. The selection of tool steel for the punch and
die and proper heat treatment will also help prevent galling. Both the punch and die
should be made from tool steel alloyed to give maximum toughness. An AISI type A6
(air-tough) or type L6 (oil-tough) tool steel is probably the best selection, especially if
large lateral forces will be encountered.
It is best to feed the work strip the long way whenever possible, especially if the blank
is to be cut from the strip before forming, as shown in Fig. 10.12. In this case, feeding
problems would be encountered, as blanks fed the short way have a tendency to twist
and bend. Feeding the long way can be simplified by using an extended stock guide
attached to the die block to guide the strip (see Fig. 10.13). A stock guide of this type
also enables the operator to feed the strip without getting his hands too close to the
die. Extended stock guided are usually not necessary in progressive dies because
previous stations serve as a guide, as shown in Fig. 10.14.
690 Tool Design

Fig. 10.12 Feeding strip: (a) Material fed short way. Loose blanks difficult to feed into form-
ing station (b) Material fed long way. Blank is fed over forming station before
cut-off

Fig. 10.13 Simple solid form die

Fig. 10.14 Progressive form die for workpiece shown in Fig. 10.13
Design of Sheet–Metal Bending, Forming and Drawing Dies 691

Consideration must be given to removing the workpiece after it is formed. The formed
workpiece will either remain in the die cavity or will cling to the punch after forming,
depending upon the shape of the workpiece. Stripper or ejector pins may be used to
remove the part from the die cavity as shown in Fig. 10.13. The stripper pins may be
knurled on the end to prevent the workpiece from the punch.
In many instances, the use of an air blast through a port in the die or the end of the
punch is an entire and simpler way of removing formed workpieces. An additional air
blast to remove the workpiece from the die set may also be used. The blast should be
timed to the press crank in order to come at the most advantageous time.

Pressure-pad forming dies A pressure pad may be necessary when the workpiece
contains sharp radii or intricate detail or when the workpiece may shift during the
forming operation. Pressure pads are applied above or below the workpiece, depend-
ing upon the design of the die and type of bend. Pressure for the pad may be applied
by springs, air, or hydraulics. Springs have the disadvantage of increasing pressure
with extended travel. Each increment of travel increases the pressure on the pad. This
may cause excessive stretching and tearing of the metal it it is necessary to move
between the pad and punch face. Short springs may be located directly under the
pad, as shown in Fig. 10.15a. Longer travel distances may require end springs to be
mounted externally under the die set or bolster plate (see Fig. 10.15b).
Pad pressure by pneumatics is provided through the use of air cushions, consisting
basically of an air cylinder, a piston, and a surge tank. The air cushion is attached
below the bolster plate or die shoe and in some applications may be attached to the
692 Tool Design

Fig. 10.15 Spring-type pressure pads

frame of the press (see Fig. 10.16). Either


the piston or the cylinder may be fixed,
depending upon the application. The pad
may be attached directly to the piston rod,
or pressure may be transmitted by attach-
ing a pressure plate to the rod which con-
tacts sliding pressure pins.
In operation, the air in the cylinder is com-
pressed on the down stroke and forced
to flow into the surge tank. On the return
stroke, the air in the surge tank forces
the cylinder back to its original position.
Pneumatic air cushions generally exert
pressure of 0.7 MPa psi or less and are
therefore limited to operations that do not
require high pressure. They are excellent
when fast operating strokes are necessary.
Fig. 10.16 Principle of air cushion
They hold a more constant pressure over
large distances of travel, and the pressure may be varied within limits. One disadvan-
tage is that the pad must be located in the die shoe because of space limitations in
the punch shoe.
Air cushions may also be used for stripping and drawing operations. They are also
used when there is danger of overloads that would damage expensive dies. The
capacity of die cushions for all operations is roughly 15 to 20 per cent of the rated
capacity of the press, with a stroke length of one-half the length of the press stroke.
Hydropneumatic die cushions may be used to overcome some of the limitations of
pneumatic die cushions. Figure 10.17 shows a hydropneumatic die cushion mounted
on an OBI press. Pressure pins are used to transfer the pressure from the cushion to
the pad or other die member to which pressure is to be applied. An installation of this
Design of Sheet–Metal Bending, Forming and Drawing Dies 693

type is simple but may serve many applications, only mirror adaptive changes from
one to the other being required. A unit of this type is more compact than a comparable
pneumatic cushion.

Fig. 10.17 OBI press equipped with a hydropneumatic die cushion (Die-Draulic, Inc.)

The hydropneumatic principle is also applied to small pistons used to replace springs.
Figure 10.18 shows this type of application and the components and mechanics. The
area below the pistons is filled with oil from the reservoir tank at air-line pressure of
around 0.7 MPa. The oil passes from the reservoir tank through the flexible high-
pressure hose and passageway drilled into the die. When the die closes, the pressure
pad is forced against the pistons, causing pressure against the oil. The pressure is
transmitted through the drilled passageway and high-pressure hose. The check valve
prevents the oil from flowing through the original passage from the tank. The oil is
forced to flow through the control valve, which restricts oil flow and therefore permits
pressure adjustment. During the downward movement of the press the pressure-
sensitive control valve maintains the pressure at which it has been set.
The advantages of this system are as follows:
1. Full pad pressure from the start
2. Constant pressure throughout the stroke of the press
3. Small pistons that will fit spring holes
4. Pressures to 310 MPa
5. Easy adjustment of pressure, eliminating changing springs for more or less
pressure
6. Elimination of pumps and motors
694 Tool Design

Fig. 10.18 Components and mechanics of a hydropneumatic die cushion


(Die-Draulic, Inc.)

Figure 10.19 shows a typical forming die equipped with hydrpnemuatic pressure pads.
The workpiece is a tray and is completely shaped in one stroke at a single station. The
four shaping operations are vertically progressive (blank, form, draw, curl) with the
curling operation being completed on the upstroke of the press.

Fig. 10.19 Die to blank, form, draw and curl (on upstroke) in one stroke
of press (Die-Draulic, Inc.)

Curling dies A curling die rolls a raw edge of sheet metal into a roll, or curl, as
shown in Fig. 10.20. The purpose is to strengthen the raw edge, provide a protective
edge, and improve the appearance of the product. The curl is often applied over a wire
ring for increased strength.
Design of Sheet–Metal Bending, Forming and Drawing Dies 695

The sheet metal to be curled must be soft


enough to roll properly. Metal that is tempered
tends to produce flats instead of round rolls.
Metal that does not respond properly to curl-
ing dies may be provided with a slight bend on
the edge in the direction of rolling to help start
the curl. The starting radius may be formed
during the previous blanking operation. Care
should be taken to ensure that the burr side of
the blank is on the inside edge of the curl.
Curling dies have a tendency to gall and for
this reason should be smoothly polished, prop- Fig. 10.20 Principle of curling die
erly heat-treated, and free from toolmarks. A
lubricant should be used during the curling operation. Once metal is transferred to the
die surface, the raw sheet-metal edge tends to “back up” and results in uneven and
defective curls.
The size of the curling groove in the die is the same as the curl diameter on the part.
The size of the curl is determined by the metal thickness. Generally, it should not
have a diameter of less than four times, the metal thickness. When a groove in the
die is larger than the curl diameter it produces, the groove diameter is too large, in
other words, when the groove diameter is too large, the curl tends to form its own
diameter.
Figure 10.19 shows a vertically progressive die which contains a curling station in one
of the four metal-shaping stations.

Embossing dies Embossing is shallow forming operation in which the workpiece


material is stretched over a male die and caused to conform to the male-die surface
by a mating female-die surface. In other words, the workpiece material is displaced
between a male and female surface. The finished product will have a depressed detail
on one side and a raised detail on the other. The major difference between embossing
and forming is that the displaced pattern is much smaller and shallower. An embossed
pattern may have more intricate detail than a formed pattern.
Embossing is used to stiffen and strengthen a sheet-metal part or to impart a raised
or depressed design on the surface of the part. The circular grooves on the bottom of
a sheet-metal container are a good example of embossing for stiffness and strength.
The metal military button with insignia raised on the surface is an example of emboss-
ing for detail.
Embossing simple shapes such as stiffening ribs does not require that the female
die to be exact reverse of the male punch (see Fig. 10.21a). Here the metal stretches
over the punch and across the radius edges of the die face. Care should be taken to
provide radii on surface over which the metal must flow.
Detail containing patterns that are reversed in both directions requires that the female
die be the exact reverse of the male die, with allowance made for the thickness of the
metal (see Fig. 10.21b).
696 Tool Design

Fig. 10.21 Embossing dies

It may be necessary to provide pressure pads or ejector pins to remove the embossed
pins to remove the embossed part from the die, depending upon the size and shape
of the embossed form.
Embossing dies are limited to rather shallow detail. If the metal tears or wrinkles with
a particular design, the use of a drawing quality or more ductile material may help. The
problem areas within the die may also be toned down or made with less relief.
Figure 10.37 shows an embossing operation built into a vertically progressive die.

Coining dies Coining is the process of pressing material in a die so that it flows
into the spaces in the detail of the die face as shown in Fig. 10.22. Coining differs
from embossing in that in coining the metal flows, whereas in embossing the metal
does not change in thickness to any great extent. Coining operations are generally
performed cold, and for the most part all surface of the workpiece and confined or
restrained.

Fig. 10.22 Coining dies


Design of Sheet–Metal Bending, Forming and Drawing Dies 697

The metal slug to be coined should be nearly the size of the die cavity and contact the
die surface over a large area because only a small amount of metal will flow during a
single hit of the press. When it is necessary to redistribute a large amount of metal, a
preliminary, metal-redistribution process such a forging or extrusion should precede
the coining.
Coining has two major advantages: (i) ornate detail can be reproduced with excel-
lent surface finish, and (ii) very close tolerance can be held. Metal buttons tableware,
medallions, medals, coins, jewelry and decoration items are example of workpiece
produced by coining when ornate detail is required. Closed tolerance can be held
because of the inherent dimensional accuracy of precision coining dies. The accuracy
produced by carefully designed and constructed coining dies is equal to the very best
machining practises and can be obtained at one hit of the press. Many semifinished
small workpiece are final-sized by coining.
Drop hammers of knuckle-joint and hydraulic presses are generally used for coining
operation because of the heavy forces required; however, coining may be done satis-
factorily on any type of press providing it has the necessary capacity and precautions
are taken to prevent overloading. For example, the press capacity for coining the
American dime is approximately 35 tons. Coining pressure is high in order to squeeze
the material into the die detail, and to accomplish this it must be from three to five
times the compressive yield strength of the material. Extra caution must be taken to
ensure that the dies are strong enough to withstand the severe pressure. Heavy die
sections are necessary, and screws and dowels must be located well away from the
die cavity.
When die sets are used for holding die blocks, they should be as strong as possible
(use steel die sets). It is advisable to specify extra-thick die shoes and punch holders
even if they must be obtained by special order. When standard die sets are used, they
can be strengthened by heat-treated tool-steel backup plates between the coining
dies and die set components.
As stated before, metal to be coined should be contained as much as possible. For
small items, such as medallions or coins, the die is contracted as shown in Fig. 10.22b
and c. Larger parts, such as silverware handles or workpieces that have only a small
section coined for accuracy, may not be totally contained in the die. Only the coined
section is enclosed by the die. The thickness of the coined workpiece is controlled
by the die block, as shown. Precision coining dies have no flash, however, flash is
allowed in a few applications and must be trimmed in a secondary operation.
One of the reasons for the coining process is to smart is to impart god surface finishes,
and therefore the surface of the coining dies must be highly polished and free from
scratches and toolmarks. The slightest scratch will transfer onto the workpiece
surface. Scratches and marks may also provide stress-concentration points that will
cause the die to split when heavy coining pressure is applied.

Bulging dies Bulging is an internal forming operation used to expand portions of


a drawn shell or tube. The forming force is applied from inside the workpiece and is
transmitted through a medium that will flow and not compress. The more common
media are rubber, urethane, heavy grease, oil, or water (see Fig. 10.23).
698 Tool Design

Fig. 10.23 Urethane bulging die: (a) Split horizontally (b) Split vertically

A bulging die must be cut into sections that open, in order to remove the workpiece
after bulging. Generally, two halves are satisfactory. The die halves must be positively
retained during the bulging operation and be equipped with quick-acting clamps or
latches. Bulging operations using liquid as a pressure-dispersing medium must be
carefully made to keep leakage to a minimum.
Rubber or urethane is preferred as a pressure dispersing medium because they are
clean and easy to use. Rubber must be of die quality and may be purchased from
die-accessory companies in cylinder or square pads. Urethane, a plastic material with
characteristics of rubber, is replacing rubber in many instances because it is more
resistant to abrasion, tears and cuts and impervious to oil and grease. Urethane may
also be purchased in standard pads, bars, and cylinders of various durometers (hard-
ness). Nonstocked sizes and durometers are available on special order.
Figure 10.23a and b shows a typical urethane bulging die. The urethane insert has
been machined to the correct size and is inserted into the workpiece by hand or held
by a stripper bolt, as shown, depending upon the workpiece design.
There are limits to the amount of deflection of the urethane material, depending upon
its hardness. Generally, urethane of 80 durometer can be deflected up to 40 per cent,
while 90 durometer can be deflected up to 30 per cent and 95 durometer up to
Design of Sheet–Metal Bending, Forming and Drawing Dies 699

20 per cent. The lower-hardness urethane is the best selection when smooth shapes
are formed. It may be necessary to use harder urethane when there are sharp angles
and detail in the form. Once the design has been tried and proved, the hardness of
the urethane should be recorded for easy duplication.
Grease, oil, water and other liquid
media are used only when the shape
of the workpiece prevents the use of
rubber or urethane (see Fig. 10.24).
Liquid or semiliquid media pose a
handling problem and may be quiet
messy. The punch and die must be
designed to confine the medium
securely in the desired area. The
press must be fast-acting to prevent
the medium from escaping between
the punch and die before the bulging Fig. 10.24 Bulging die using liquid medium
operation is completed.
In Fig. 10.24, the workpiece is filled with the liquid medium while it is locked in position
between the dies. The punch forces the medium into the workpiece, which expands at
the die cavity. A bulging die of this type is used only for limited production because of
the time consumed in producing the part.

Assembly dies In this type of tooling, two or more parts are assembled, held in posi-
tion, and then locked in position by riveting, staking, crimping, or press fitting. The
operator may physically place the parts in position in the die, as shown in Fig. 10.25a.
Here two studs are riveted to a strap. A large quantity of parts may justify a more
complicated die that would automatically assemble the parts in the die while being
fed from a hopper.
Various forming operations used primarily for assembly purposes are shown in
Fig. 10.25. Staking is shown in b and c. Tab staking consists of inserting precut tabs
through slots in mating parts and forming over the tabs to hold the assembly together.
The upset stake is used primarily to secure machined parts to sheet metal. The
machined part is upset in several places along the joint to form retaining lips.
Crimping is similar to staking, except that sheet metal is folded around the matting
workpiece. Electrical connections are commonly attached to wires by crimping.

10.4 DRAWING OPERATIONS


Drawing is the name given to the production of cups, shells, boxes and similar articles
from metal blanks. A simple drawing operation is shown in Fig. 10.26. A round blank
is first cut from flat stock. The blank is then placed in the draw die, where the punch
pushes the blank through the die. On the return stroke, the cup is stripped form the
punch by the counterbore in the bottom of the die. The top edge of the shell expands
(springback) slightly in order to make this possible. Note that the amount of spring-
back in the drawing has been exaggerated. The punch has an air vent to prevent a
vacuum from being forced when the part is stripped from the punch.
700 Tool Design

Fig. 10.25 Forming operations used primarily for assembly purpose: (a) Riveting (b) Tab
stake (c) Upset stake (d) Crimping

Fig. 10.26 Simple drawing operation

The die previously described is limited to shallow cups. Deeper draws must have a
provision for confining the metal in order to prevent excessive wrinkling and other
Design of Sheet–Metal Bending, Forming and Drawing Dies 701

defects. A rigid blankholder is generally used, as shown in Fig. 10.27a, on heavier


gauge materials that have only a moderate tendency to wrinkle. The spring-loaded
pressure pad or the pressure pad shown in Fig. 10.27b is more commonly used than
the rigid blankholder. Here the punch pushes the blank through the die, and the spring-
loaded pressure pad extends over the nest, which acts as a spacer and is machined
to a thickness that will provide the proper pressure on the blank at all times. In other
words, the pressure on the workpiece is even during the complete drawing cycle and
does not increase, as would be the case if the spacer were not present.

Fig. 10.27 Methods of retaining metal in draw dies: (a) With rigid blank holder (b) With
spring pressure pad

Generally speaking, a drawing operation is referred to as shallow drawing when the


cup is no deeper than half its diameter. There would be vary little thinning of the
metal in this case. Deep drawing is drawing a cup deeper than half its diameter.
Designing dies for deep drawing requires considerable skill and experience, and only
the general principles can be studied here. Extreme drawing as shown in Fig. 10.28,
is possible only by dividing the work into several stages or draws. Here a blanked
702 Tool Design

part to drawn into a shell through successive first, second, third and fourth drawing
operations. Deep drawing is often accompanied by an operation called ironing, which
is defined as the reduction in wall thickness of a shell by forcing it through a tight die.
The thickness of the shell bottom is not changed during an ironing operation, but the
walls are both lengthened and made thinner.

Fig. 10.28 Stages in deep drawing (E W Bliss Co.)

Drawn parts are commonly of circular form in a transverse section perpendicular to


the axis of the punch. Rectangular and other shapes may also be produced, but they
are usually more difficult because metal distortion is not even. Drawn shells may have
tapered sides as well as straight sides. Most ductile metals can be drawn, although
drawing practise may vary somewhat for different work materials.

Metal flow during drawing To understand what happens to the metal during a
drawing operation, the process should be broken into progressive stages of formation
as shown in Fig. 10.29. Here a flat circular blank is drawn into a flat bottomed cup by
forcing a punch against the blank, which rests on a die. Note that the blank is divided
into sections for easy reference. During the first stage, the punch contacts the blanks
as shown in b. The area of punch contact is denoted as section 1. This section forms
the flat bottom of the cup and is not distorted by the punch. Upon further penetration
of the punch, the metal in section 2 is bent or wrapped around the punch nose and
die radius. As shown in c.
As the punch penetrates still further, the metal that was previously bent over the die
radius becomes straightened. Additional metal is pulled over the die radius, which in
time is straightened as the punch progresses downward. In other words, the outside
edge of the blank is being drawn toward the punch. In order to accomplish this suc-
cessfully, the section of the blank must reduce in circumference the only way to do
this is for the metal to compress and become thicker. The amount of compression and
thickness increase as the blank edge nears the punch. The metal may tend to wrinkle
rather than compress, especially in thin sheets or with deeper draws. A blank holder is
used to prevent the formation of wrinkles in this case. There must be enough force on
the blank holder to prevent formation of wrinkles in this case. There must be enough
force on the blank holder to prevent formation of wrinkles because after a wrinkle
is started, the blank holder is raised from the surface of the metal and allows other
wrinkling varies from practically zero in the case of the fixed blank holder to approxi-
mately one-third the drawing pressure. Blank-holder pressure is generally determined
by past experience and trial and error because of the many variables that are involved
in the drawing operation.
The force created by the blank holder also increase frictional forces. Too much blank-
holder pressure may tear the sidewall of the drawn cup. The sidewall transmits the
punch force to the area of draw on the die radius. Tearing will generally occur near the
punch radius because this area will thin due to tension forces.
Design of Sheet–Metal Bending, Forming and Drawing Dies 703

Fig. 10.29 Metal flow during drawing

10.5 VARIABLES THAT AFFECT METAL FLOW


DURING DRAWING
The many variable in the drawing are too numerous to mention in this text, but a few
of the important ones are as follows.

Radius on punch There is no set rule for the size of the radius on the punch. A
sharper radius will require higher forces when the metal is folded around the punch
nosed and may result in excessive thinning or tearing at the bottom of the cup. A gen-
eral rule to prevent excessive thinning is to design the punch with a radius of form 4 to
10 times the metal thickness. If the radius must be less than four times the metal thick-
ness, it may be necessary to form it over a larger punch radius and then restrike it to
704 Tool Design

develop the specify radius. The radius may be determined by the product design when
only one draw is necessary to complete the workpiece because the cup bottom takes
the shape of the punch nose. When several redraws are required, the punch radius for
each redraw should be proportionately smaller than that of the preceding shell.

Draw radius on die Theoretically, the radius on the draw die (draw ring) should be
as large as possible to permit full freedom of metal flow as it passes over the radius.
The draw ring causes the metal to being flowing plastically and aids in compressing
and thickening the outer portion of the blank. However, if the draw radius is too large,
the metal will be released by the blank holder too soon and wrinkling will result. Too
sharp a radius will hinder the normal flow of the metal and cause uneven thinning of
the cup wall, with resultant tearing. The general rule is to make the draw radius four
times the material thickness. The draw radius may be increased to six to eight times
the metal thickness when drawing shallow cups of heavy-gauge metals without a
blank holder. The nomograph in Fig. 10.30 gives a more exact method of determining
draw-die radius, based on the relationship of the blank diameter to the cup diameter.

Example:
d = 25 mm, h = 20 mm, T = 0.8 mm
______ ____________
D = ÷d2 + 4dh = ÷252 + 4 × 25 × 20 = 51.23 mm
D – d = 26.23 mm
R = 3.66 ± 0.125 mm

Fig. 10.30 Nomograph for determining draw-die radius

The relationship of the punch-nose and draw-die radii to minimise stock thinning is
shown in Fig. 10.31. The centre point of the draw radius should be approximately 3.2
Design of Sheet–Metal Bending, Forming and Drawing Dies 705

mm outside the previous cup, as illustrated in


A. The centre point of the punch-nose radius
should be slightly inside the following shell as
in B. The centre points of the punch-nose radii
on the last two operations are about on the
same line, thereby maintaining the falt on the
bottom of the cup as in c.

Friction The force of static friction between


the workpiece blank and draw-die surfaces
must be overcome in a drawing operation. The
force of the blank holder adds significantly to
the force of static friction. Once static friction is Fig. 10.31 Relation to punch-nose
overcome by the start of the blank movement, and die-draw radii. [From
continuous movement of the punch is quite ASTME, “ASTME Die
important because the force needed to over- Design Handbook,” FW
come dynamic friction is less than that needed Wilson (ed.), McGraw-
for static friction. Hill, New York, 1965, by
permission.]
Since the blank-holder pressure causes higher
frictional forces, it should be only high enough to prevent wrinkling of the metal. Blank-
holder pressure that is too high will cause the metal to be restricted and result in tear-
ing of the cup wall.
Blank-holder pressure cannot be reduced below the point where wrinkling of the
metal occurs. A lubricant is generally applied to reduce friction. Shallow draws in light
stampings can be produced with little or no lubrication, but when forces become larger
and scoring, wrinkling, and tearing become a problem, a lubricant is almost always
used.
The purpose of a press lubricant is to provide a film between the workpiece and the
punch and die. The film must be strong enough to permit metal deformation without
being squeezed from the surfaces. Table 10.1 gives various compositions for press
lubricants for stamping and deep drawing. The following paragraphs give a brief expla-
nation of Table 10.1. The subsection on cutting fluids in Chapter 4 should be reviewed
if the student is not familiar with basic lubricant terms.
When pressures are low, straight mineral oil, general-purpose soluble oil, or a
diluted soap solution can give satisfactory service, since pressures will not rupture
the lubricating film. As pressure become greater, lubricants containing sufficient oili-
ness (polar materials such as fatty oils, waxes and concentrated soaps) are required.
These physically adherent materials are adsorbed to metal surfaces, maintaining a
persistent microscopically thin film where lubricants lacking sufficient oiliness would
be squeezed out. When pressure is very high, as in a severe draw, the corresponding
rise in temperature reduces the adhering of the lubricant.
Some form of extreme-pressure lubricant is then required if welding (galling or seizure)
between the tool and work surface is to be avoided. Welding is evidenced by metal
build-up on the punch and die, causing scratch marks on the workpiece. Improper
lubrication may also result in tearing the metal, wrinkling, buckling, and other causes
for rejection of the work.
Table 10.1 Press-Lubricant composition for stamping and deep drawing with comparison of properties*
Composition Capability of Cleanability by Rust Non ferrous metal staining Remarks
prolonging Vapor Aqueous Protection
die life degreasing method
706 Tool Design

Chlorinated Oils:
Concentrate (non-emulsifiable) 1 3 5 5 Usually none Light colour, almost no odour
Concentrate (emulsifiable) 1 3 2 5
Mineral-oil bend, high chlorine
content, emulsifiable 1 2 1 5
Mineral-oil bend, low chlorine
content, emulsifiable 2 1 1 5
Sulphurised oils:
Concentrate 2* 3 4 3 Nonstaining to aluminium, zinc Brown colour; typical sulphur
Mineral-oil bend, high sulphur content 4* 2 3 3 and tin; active sulfur cause odour
Low sulphur content 5* 1 2 3 black stain on copper, brass
and bronze
Fatty oils and fats:
Concentrate 2* 3 3 3 Usually none; copper alloys Typical fatty odour; burns
Mineral-oil blend 4* 1 2 3 develop green stain with off above 1000°F leaving
compounds having high free no residue
fatty-acid content
Soap-fat compounds:
Nonpigmented 3 4 3 3 Usually tarnishes brass, bronze,
Pigmented 2 5 4 2 zinc, and tin
Liquid lubricating soaps 3 5 1 3 Usually tarnishes brass, bronze, Foams
zinc, and tin
Contd...
Contd...
Composition Capability of Cleanability by Rust Non ferrous metal staining Remarks
prolonging Vapor Aqueous Protection
die life degreasing method
Dry-film soaps 1 5 2 2 Usually tarnishes brass, bronze, Requires special equipment;
zinc, and tin best suited for production
work on larger parts
Mineral oil 5 1 2 3 None Wide range of viscosities
Soluble oils:
General-purpose 5 2 1 3 May tarnish brass, bronze, zinc
Heavy-duty 3 2 1 3 and tin
Vanishing oil 5 … … … None Evaporates from metal with
in 1 h exposure to air leaves
no residue.
* Ratings: 1. best; 2. good; 3. average; 4. poor; 5. worst.
† Rating indicated is for ferrous metals. Ration would be upgraded for nonferrous metals.
SOURCE: Tooling and Production Magazine
Design of Sheet–Metal Bending, Forming and Drawing Dies 707
708 Tool Design

The available Extreme Pressure (EP) agents function either chemically or mechani-
cally. Those which provide EP characteristics chemically usually contain loosely com-
bined chlorine or sulphur, which reacts with the punch, die, and work to form chemical
protection films that are highly resistant to welding. Lubricants that function mechani-
cally minimise friction by incorporating powdered spacing agents, i.e., pigments, such
as chalk, graphite, or molybdenum disulphide. These substances act as physical
separators between the tools and the work.
Pressworking lubricants may be applied by roller-coating, brushing, swabbing, spray-
ing, or flushing. The method depends largely on the viscosity of the lubricant. The
lubricant is removed from the workpiece with an alkaline wash, with emulsifiers, or by
vapour or solvent degreasing, depending upon the type of lubricant.
Other factors influencing friction are the finish on both sides of the work material and
the surface finish on the punch, die, and blank holder. In other words, a smoother sur-
face on the work material and the mating die surface will result in less friction.

Material to be drawn The characteristics of the material to be drawn have a great


influence on the success of a drawing operation. Ductility and yield strength are the
most important. Ductility is the ability of the metal to undergo a change of shape occur
in a metal. A low yield strength is desirable so that metal flow can beign easily without
tearing near the punch radius. Ductility and yield strength are generally determined
by the product designer when he specifies the material, and the tool designer must
work around these variable. However, it is sometimes possible to request a change to
a material with better drawing properties for complicated shapes; therefore, the tool
designer must have a good understanding of the material properties.
The following principal factors affect the selection of grade and quality of low
-carbon-steel for deep drawing:1
1. Severity of draw as determined by the amount of reduction and punch-nose
radius.
2. Thickness of sheet
3. Shape of part (round, rectangular, or conical)
4. Flange requirements
5. Ironing requirements
6. Desired finish
7. Grain size
8. Press speed
9. Availability of material
10. Cost
The three principal classes of available low-carbon sheet steel are commercial
quality, drawing quality, and aluminium-killed (special-killed) in order of increasing
drawability.
For hot-rolled sheets, the mill oxide on the surface is detrimental for deep drawing
because it impairs the flow of metal and increase tool wear. Special surface can be
obtained for any of the three qualities at extra cost, or oxide can be removed by

1
ASM, “Metals Handbook,”8th ed., vol. 4, “Forming,” Metals Park, Ohio, 1969.
Design of Sheet–Metal Bending, Forming and Drawing Dies 709

pickling, which also adds to the cost. Hot-rolled sheets for deep drawing are usually
purchased pickled and oiled.
The primary consideration in choosing between hot-and cold-rolled steel for deep
drawing is the availability of the required sheet thickness. Where superior surface fin-
ish is required, cold-rolled sheet or strip is specified. The choice between commercial
or drawing quality in either the hot- or cold-rolled sheets depends upon the severity
of the draw.
Drawing quality is generally for severe draws because, with a nominal extra cost,
required drawability is guaranteed. The problem is to recognise when it is expedient to
pay the extra price for drawing performance guaranteed by the supplier.
Rimmed steels are satisfactory for most draws of mild to moderate severity. The use
of special-killed steel is usually limited to cold-rolled steel for draws requiring freedom
from strain aging, maximum uniformity within a shipment and among successive ship-
ments, and maximum drawability.
The first approach to selection can be made as follows:
1. When surface conditions of the finished piece are most important, a cold-rolled,
rimmed steel will be specified.
2. When drawability requires freedom from aging, or if the design specifies severe
forming, such as flanges, ironing, sharp corner radii, extreme cup depth, or
lack of symmetry, special-killed quality is indicated.
Assuming correct tooling and normal reductions, any scrap losses with special-killed
steel will probably result from the opening of small surface disturbances during draw-
ing. This occurs more often in coil stock than sheet.
Special soundness refers to the internal condition and is not a prerequisite for deep
drawing. It is specified only when subsequent operations such as bright plating,
highly critical welding, or restrictive carburising require maximum uniformity and
homogeneity.
Table 10.2 shows the grades of low-carbon steel recommended for drawing round
cups. Severity is based on cup reduction and punch-nose radius for specific sheet
thicknesses. Optimum tooling and lubrication are assumed for the most difficult draws.
Similarly, Table 10.3 shows grades recommended for drawing rectangular cups or
boxes having corner radii equal to the punch-nose radius. In both tables, severity of
drawing increase toward the upper right corner of each of the subdivisions that define
the various conditions, The sequence of steels recommended to meet the increasing
severity conditions is commercial quality, drawing quality, and drawing quality special-
killed.

Per cent reduction and depth of draw The per cent reduction in drawing cylindrical
shells is generally expressed in terms of the diameters of the blank D and the drawn
shell d, where D equals the OD of the blank and d equals the ID of the shell. This
percentage provides an approximate value for the amount the work material is to be
compressed. The drawability of a metal is often expressed as the percentage reduc-
tion from the blank diameter to the cup diameter. Percentage reduction is calculated
from the formula

(
P = 100 1 – __
D
d
)
Table 10.2 Selection table for low-carbon sheet for deep drawing of cylindrical cups*
710 Tool Design

Two entries in one space indicate that the choice is borderline and will be influenced by uniformity, price, quantity, and availability, omission of an entry indicates that
the draw is usually too severe to be completed successfully.

Reduction in drawing†
Punch-
nose 1 mm thick 2.7 mm thick 6.4 mm think
radius

20% 30% 40% 50% 20% 30% 40% 50% 20% 30% 40% 50%

No flange, no ironing
2t CR, DQ
4t CR, DQ CR, DQ CR, DQ, SK HR, DQ HR, DQ HR, DQ HR, CQ HR, CQ HR, DQ
CR, CQ CR, DQ
8t CR, CQ CR, CQ CR, DQ CR, DQ, SK HR, CQ HR, DQ HR, DQ HR, DQ HR, CQ HR, CQ HR, DQ
CR, CQ CR, CQ
16t CR, CQ CR, CQ CR, CQ CR, DQ HR, CQ HR, CQ HR, CQ HR, DQ HR, CQ HR, CQ HR, CQ HR, DQ
CR, CQ
2t CR, CQ CR, CQ CR, CQ CR, DQ HR, CQ HR, CQ HR, CQ HR, CQ HR, CQ HR, CQ HR, CQ HR, DQ
No flange, ironed‡
2t CR, DQ, SK
4t CR, DQ CR, DQ, SK HR, DQ HR, DQ HR, DQ, HR, DQ HR, DQ
CR, DQ
Contd...
Contd...
Reduction in drawing†
Punch-
nose 1 mm thick 2.7 mm thick 6.4 mm think
radius
8t CR, DQ CR, DQ, SK CR, DQ, SK HR, CQ HR, CQ HR, DQ HR, DQ HR, CQ HR, DQ HR, DQ
CR, DQ
16t CR, DQ CR, DQ CR, DQ, SK CR, DQ, SK HR, CQ HR, CQ HR, DQ HR, DQ HR, CQ HR, CQ HR, DQ HR, DQ
32t CR, DQ CR, DQ CR, DQ CR, DQ, SK HR, CQ HR, CQ HR, CQ HR, DQ HR, CQ HR, CQ HR, CQ HR, DQ
With flange, no ironing
2t
4t CR, DQ CR, DQ, SK CR, DQ, SK CR, DQ CR, DQ, SK CR, DQ, SK HR, DQ HR, DQ
8t CR, DQ CR, DQ CR, DQ, SK CR, DQ, SK HR, DQ CR, DQ CR, DQ CR, DQ, SK HR, DQ HR, DQ HR, DQ HR, DQ
16t CR, CQ CR, CQ CR, DQ CR, DQ HR, CQ HR, CQ, HR, DQ, CR, DQ HR, DQ HR, DQ HR, DQ HR, DQ
CR, CQ CR, DQ
32t CR, CQ CR, CQ CR, CQ CR, DQ HR, CQ HR, CQ HR, DQ, CR, CQ HR, CQ HR, DQ HR, DQ HR, DQ
CR, CQ
With flange, ironed‡
2t
4t CR, DQ, SK CR, DQ, SK CR, DQ, SK CR, DQ, SK CR, DQ, SK HR,DQ
8t CR, DQ, SK CR, DQ HR, DQ CR, DQ CR, DQ, SK HR, DQ HR, DQ
16t CR, DQ CR, DQ CR, DQ, SK CR, DQ, SK HR, DQ HR, DQ HR, DQ CR, DQ, SK HR, DQ HR, DQ HR, DQ HR, DQ
CR, DQ CR, DQ
32t CR, CQ CR, CQ CR, DQ CR, DQ, SK HR, DR HR, DQ HR, DQ HR, DQ HR, DQ HR, DQ HR, DQ HR, DQ
CR, DQ
*CR = cold-rolled; CQ = commercial quality; DQ = drawing quality; HR = hot-rolled; SK = special-killed
† In terms of reduction from blank diameter to cup diameter.
‡ Operation that will bring the sidewall to a uniform thickness with the minimum amount of reduction.
Design of Sheet–Metal Bending, Forming and Drawing Dies 711

SOURCE: ASM, “Metals Handbook”, Vol. 4. “Forming,” Metals Park, Ohio, 1969.
Table 10.3 Selection table for low-carbon sheet for deep drawing of rectangular box-shaped cups without a flange*
712 Tool Design

Two entries in one space indicate that the choice will be influenced by uniformity, price quantity and availability; omission of an entire indicates that the draw is too
severe to be completes successfully.

Punch- Reduction in drawing†


nose 0.036 in. thick (20 gauge) 0.105 in. thick (12 gauge) 0.250 in. thick
radius 20% 30% 40% 50% 20% 30% 40% 50% 20% 30% 40% 50%

No flange, no ironing
4t
8t CR, DQ, SK HR, DQ HR, DQ HR, CQ HR, DQ
CR, DQ
16t CR, DQ, SK CR, DQ, SK CR, DQ, SK HR, DQ HR, DQ HR, DQ HR, CQ HR, DQ HR, DQ
CR, DQ
32t CR, DQ CR, DQ, SK CR, DQ, SK HR, DQ HR, DQ HR, DQ CR, DQ, SK HR, CQ HR, CQ HR, CQ HR, DQ
CR, DQ
No flange, ironed*
4t
8t CR, DQ, SK HR, DQ CR, DQ HR, DQ
16t CR, DQ, SK CR, DQ, SK HR, DQ HR, DQ CR, DQ, SK HR, CQ HR, DQ HR, DQ
32t CR, DQ CR, DQ, SK CR, DQ, SK HR, DQ HR, DQ HR, DQ CR, DQ, SK HR, CQ HR, CQ HR, DQ HR, DQ
CR,DQ
* CR = cold-rolled; CQ = commercial quality; DQ = drawing quality; HR = hot-rolled; SK = special-killed
† In terms of reduction from blank width to cup width. Corner radii are the same as punch nose radius.
‡ Operation that will bring the sidewall to a uniform thickness with the minimum amount of reduction.
SOURCE: ASM, “Metals Handbook”, 8th ed., Vol. 4. “Forming,” Metals Park, Ohio, 1969.
Design of Sheet–Metal Bending, Forming and Drawing Dies 713

where P = percentage reduction


d = ID of drawn shell
D = OD of blank
When using sheet-metal drawing formulas, there is often considerable confusion
whether d equals the inside diameter or the outside diameter of the drawn cup. The
per cent-reduction formula is based on the relationship between the punch diameter
and the blank diameter; therefore, it would theoretically equal the inside diameter of
the drawn shell. However, the change in the answer is negligible when using d as
equal to the outside diameter, especially when the wall is thin. For the sake of consis-
tency throughout the examples given in this section, d will equal the outside diameter
of the drawn shell (realising, of course, that it is not theoretically correct for the per
cent-reduction formula). This prevents confusion, since d is generally equal to the
outside shell diameter for the majority of sheet-metal drawing formulas.
The theoretical maximum per cent reduction for one draw is approximately 50%
although the figure is hard to obtain in production. For practical purpose, it is better to
figure a maximum of 40% reduction for a single draw.
The depth of draw, expressed as the ratio of height of cup h to the diameter of cup d,
is also an indicator of the degree of compression needed. When the ratio h/d exceeds
a value of 0.75 (depending upon the drawability of the material), more than one reduc-
tion will be necessary. Table 10.4 shows the number of probable reductions based on
the draw ratio.
After the number of probable reductions is determined, it is necessary to estimate the
cup size from the blank size or cup size from an earlier draw. This information can be
determined from Fig. 10.32. The student should study the following sample calcula-
tion closely.

Fig. 10.32 Chart for checking percentage reduction in drawing of cups. The inside diameter
is ordinarily used for the cup diameter.
714 Tool Design

Table 10.4 Possible number of reductions for a given ratio of shell height to diameter
Reduction %
Ratio h/d Number of reductions First draw Second draw Third draw Fourth draw
Up to 0.75 1 40
0.75-1.5 2 40 25
1.5-3.0 3 40 25 15
3.0-4.5 4 40 25 15 10

Table 10.5 Draw clearance


t = thickness of the original blank

Blank thickness, mm. First draws Redraws Sizing draw*


Up to 0.38 1.07t-1.09t 1.08t-1.1t 1.04t-1.05t
0.4-1.27 1.08t-1.1t 1.09t-1.12t 1.05t-1.06t
1.3-3.18 1.1t-1.12t 1.12t-1.14t 1.07t-1.09t
3.5 and up 1.12t-1.14t 1.15t-1.2t 1.08t-1.1t
* Used for straight-sided shells where diameter or wall thickness is important or where it is necessary to improve
the surface finish in order to reduce finishing costs.

Example 10.5 To illustrate the use of Fig. 10.32, assume the problem of deter-
mining whether a cup of 190 mm ID can be in three draws of
40, 20, and 15 per cent reduction, respectively, from 450 mm
diameter blank.
Solution: To find the diameter of the cup after the first draw, trace the line for the 450
mm diameter horizontally to its intersection with the diagonal for 40% reduction. From
this intersection, draw a vertical line to the top of the chart and read 276 mm. For the
ID of the cup after the first draw. Next, trace a horizontal line from 276 on the vertical
axis until it intersects the diagonal for 20% reduction. From this point, draw a verti-
cal line to the bottom of the chart and read 221 mm, which is the ID of the cup after
second draw. The ID of the cup after the third draw, assumed to be 15% reduction, is
found using the same procedure by drawing a line horizontally from 221 to it intersec-
tion with the diagonal for 15% reduction and from there to the bottom of the chart,
which gives a reading 189 mm. Accordingly, it is concluded that a 1900 mm ID cup
can be drawn from 460 mm diameter blank in the three assumed reductions.

It may be necessary to anneal some metals before they can be redrawn, especially
between the third and fourth draw. The necessity of annealing depends upon the
work-hardening characteristics of the material. Annealing must be carefully controlled,
or it may impair drawability.

Drawing speed The velocity at which the punch penetrates the workpiece often has
a definite effect on a drawing operation. Low carbon is normally drawn from 9 to 15
m/min under normal conditions. Nonferrous work material is drawn from 45 to 60
m/min. Generally speaking, the harder and less ductile the material, the slower the
drawing speed must be. When cracking or excessive thinning occurs, the drawing
speed must be reduced. Close control of drawing speed is not always possible, espe-
cially on mechanical presses.
Design of Sheet–Metal Bending, Forming and Drawing Dies 715

Die clearance Die clearance is the gap left between the punch and die to allow for
the flow of the work material. Generally enough clearance is left to allow for thicken-
ing of the metal. This allowance ranges from 7 to 20 per cent of the metal thickness,
depending upon the type of operation and the metal. When clearance is equal to the
metal thickness or less, ironing or burnishing of the metal will occur near the top of the
cup. A certain amount of ironing may be allowable on the more ductile metals when
uniformity of wall thickness is required. Table 10.5 shows draw clearance for various
blank thicknesses.

10.6 DETERMINING BLANK SIZE


The designer of draw dies must determine the approximate blank size of drawn work-
pieces before designing the die. In addition to determining the size of the blank to
produce the drawn shell, he is also able to determine the number of draws necessary
to produce the shell. Various methods of determining blank size have been developed
over the year, based on mathematic, graphical layouts, surface area, weight, and
volume. The method used in the following discussion is based on mathematics alone,
as it is seldom possible to figure blank size accurately because of thickening and
thinning of the wall and variations in ductility of material. The exact blank size is best
determined by trial and error using the material that is to be used in the drawing of
the workpiece. Other methods of determining blank size may be found in handbooks
on the subject.
The following equations1 can be used to calculate the blank size for cylindrical shells
of relatively thin metal. The ratio of the shell diameter to the corner radius d/r can
affect the blank diameter and should be taken into consideration.

________
÷d 2 + 4dh when d/r is 20 or more
_____________
2
÷d + 4dh – 0.5r when d/r is between 15 and 20
D = __________
÷d 2 + 4dh – r when d/r is between 10 and 15
____________________________
2
÷(d – 2r) + 4d(h – r) + 2pr(d – 0.7r) when d/r is below 10

where D = blank diameter


d = shell OD
h = shell height
r = corner radius of punch
These formulas give the theoretical blank size, which is only an approximation when
applied to actual practise. Extra metal should be added to the formula blank diameter
to provide for trimming, which is generally necessary on deeper draws to eliminate
the uneven and irregular edge on the rim of the draw cup. The extra material added
to the blank diameter is referred to as trim allowance. The necessary trim allowance
increase as the size of the drawn cup increase. The general rule of trim allowance is
to increase the blank diameter by 0.05 mm for cup diameter up to 10 mm. Therefore,
add 0.05 mm for each 10 mm of cup diameter. For example, 0.5 mm to the blank
diameter for a cup diameter of 100 mm.

1
From ASTME, “ASTME Die Design Handbook,” F W Wilson (ed.), McGraw-Hill, New York, 1965.
716 Tool Design

10.7 DRAWING FORCE


The amount of force required to shape a symmetrical cup by drawing can be
calcaulted form the following formula:

(
D
P = pdt (UTS) __ – C
d )
where P = drawing pressure
d = shell OD
D = blank diamater
t = thickness of metal
UTS = Ultimate tensile strength
C = constant to cover friction and bending (0.6 to 0.7 for ductile materials)
The formula is applicable to a symmetrical cupping operation only. The stretching
and compression of metal in irregular part is too unpredictable for figuring forces by
formula. However, the above formula may be used to some extent in determining
approximate forces for workpieces similar in configuration to a symmetrical cup.

Table 10.6 Average Mechanical Properties of Various Alloys


Metal Tensile strength MPa Yield strength MPa

SAE 1020 steel, HR 380 207


SAE 1020 steel, CD 421 352
SAE 1040 steel, HR 524 290
SAE 1040 steel, CD 586 490
SAE 1095 steel, HR 827 455
SAE 4130, HR, annealed 621 414
SAE 8640, HR, annealed 634 421
AIST type 301 stainless steel 690 241
Yellow brass, annealed 324 103
Deoxidised copper, annealed 221 69
Magnesium Alcoa grade AMC35 193-255 124-193
2S aluminium:
0 temper 90 35
H16 temper 145 138
24S aluminium:
0 temper 186 76
T3 temper 448 310
52S aluminium:
0 temper 193 90
SOURCE: ASTME, “ASTME, Die Design Handbook,” F W Wilson (ed.), McGraw-Hill, New York, 1955, by
permission.

10.8 SINGLE- AND DOUBLE-ACTION DRAW DIES


Draw dies consisting of only a punch and die are known as single-action dies. The
majority of draw dies discussed thus far have been single-action dies. The simplest
type of single-action die is shown in Fig. 10.26. In this die, a precut blank is placed
in a nest on top of the die. When the punch pushes the blank through the die on the
Design of Sheet–Metal Bending, Forming and Drawing Dies 717

return stroke, the cup is stripped


from the punch by the counterbore
in the bottom of the die. An air vent
should be provided to eliminate
suction, which would tend to make
the cup cling to the die.
Figure 10.27 shows the use of
blank holders and pressure pads
to control the metal flow and pre-
vent wrinkling. The depth of draw
is limited with spring-loaded pres-
sure pads because the pressure
is not constant and increase as
the springs are compressed on Fig. 10.33 Single-action draw die with
the downstroke. This type of die spring-loaded knockout
is usually limited to shallow draws
and light-gauge material.
When a flanged shell that cannot
pass through the die is drawn, a
knockout of kicker may be used,
as shown in Fig. 10.33. The
knockout may be spring-loaded,
as shown, or actuated by an air
cylinder or cushion. It may also
be used for shallow embossing
or to form slight reentrant curves
on the bottom of the cup. The use
of spring-loaded knockouts as
shown in limited to shallow draws
because of the necessity of short
springs and uneven pressure.
Deep draws can be made on
single-action dies by inverting the
die and utilising an air cushion,
as shown in Fig. 10.36. In this Fig. 10.34 Inverted single-action die utilising air
case the punch is mounted on the cushion to supply holdown pressure
lower shoe and the die on the upper shoe. The air cushion supplies the blank-holder
force which may be held at a constant and regulated pressure. The drawn cup is
removed from the die on the upstroke of the run by using a knockout and a knockout
bar attached to the press frame.
Figure 10.35 shows a similar arrangement utilising the hydropneumatic pistons
described earlier in this chapter. The die shown is designed to blank, emboss, draw,
and pinch-trim in one stroke of the press. An integral two-valve unit is shown with
connections to the upper and lower sections of the die. Note that the cylinders in both
the punch holder and shoe have closed ends and project through. The punch holder
and shoe are supported against parallel blocks, leaving space for tubing connections
to the cylinders.
718 Tool Design

Fig. 10.35 Die to blank, emboss, draw, and pinch-trim in one stroke
of press (Die-Draulic, Inc.)

The true double-action die is shown in Fig 10.36. Here the die is designed for a dou-
ble-action press with the blank holder fastened to the outer ram, which descends first
and grips the blank. The punch is fastened to the inner ram, which descends next to
form the part. The cup may be pushed through the die or ejected back through the die
by use of knockout actuated by an air cushion. The knockout may also be used to hold
the blank firmly against the punch nose during the drawing operation or to emboss or
form reentrant curves on the bottom of the cup.

Fig. 10.36 Typical double-action draw die mounted on double-action press


Design of Sheet–Metal Bending, Forming and Drawing Dies 719

Solved Design Problem

Problem 1 Figure 10.37 shows a symmetrical-cup workpiece with a shell height of


40 mm and a shell diameter of 50 mm the corner radius is 1.6 mm. The workpiece
material is 1020 cold-rolled steel 0.8 mm thick.
Make necessary calculations for designing the die
for this drawing operation.
Solution: Determine Size of Blank:
50
d ___
__
The ratio, = = 31.25
r 1.6
Then the formula for determining the size of blank
will be
________ __________
D = ÷d 2 + 4dh = ÷502 + 4 × 40 = 102.47
This is the theoretical blank size. Since a
smooth edge is required, it will be neces-
sary to add extra metal to provide for trim-
ming. According to the general rule of adding
0.05 mm to the blank diameter for each 10 mm of
Fig. 10.37 Draw shell used in
the cup diameter, 0.005 × 102.47 = 0.51 mm should
sample problem
be added for trimming. Thus, D would be
D = 102.47 + 0.51 = 102.98 mm ª 105 mm
Determine Percentage Reduction:
Percentage reduction is calculated from

( d
) (50
)
P = 100 1 – __ = 100 × 1 – ____ = 52.38% reduction
D 105
The theoretical maximum per cent reduction for one draw is approximately 50%.
Therefore, it is obvious that the above cup cannot be drawn in one operation.
Determine Draw Ratio:
The depth of draw is expressed as the ratio h/d is 40/50 is 0.80. When the value
exceeds 0.75, more than one reduction is necessary, as was indicated. Referring to
Table 10.4, two reductions are needed, with a 40% reduction for the first draw and
a 25% reduction for the second draw. It is now necessary to determine the cup size
from the blank size for a 40% reduction and the cup size from the previous draw for
a 25% reduction. This is accomplished by using the chart in Fig. 10.32. To find the
diameter of the cup after the first draw, trace the line from the 105 mm blank diameter
horizontally to its intersection with the diagonal 40% reduction. From this intersec-
tion, draw a vertical line to the top of the chart and read 63 mm for the ID of the cup
after first draw. Next, trace a horizontal line from 63 mm on the vertical axis until it
intersects the diagonal for a 20% reduction. From this point, draw a vertical line to
the bottom of the chart and read 47.25 mm, which is the ID after second draw. This
is below the 50 mm diameter required for the finished size of the cup. Therefore, the
first draw will be a 40% reduction to 63 mm diameter, and a second draw will be 20%
reduction to the final 50 mm diameter.
720 Tool Design

Determine Radius on Punch and Die:


The general rule for determining draw-die radius is to make the draw radius four times
the material thickness. According to this rule, the draw radius for the shell in Fig. 10.37
would be 3.2 mm for the first draw. However, the nomograph in Fig. 10.30 (using
D–d = 105 – 50 = 55) gives a radius of approximately 6.63 mm. Since it is easier to
increase the radius during the tryout of the die, the designer would probably specify
3.2 mm for the draw radius, with provisions to increase the radius if tearing or exces-
sive thinning is evident.
The general rule for a single draw is to design the punch with a radius from 4 to 10
times the metal thickness. The punch radius for the first draw will be determined by
the relationship of punch-nose and draw-die radius, as shown in Figs. 10.31 and
10.38. In order to place the centre point of the second draw radius 3.2 mm outside the
previous cup, the punch for the first draw must have a radius approximately 8.8 mm
and the draw radius for the second draw must be 12.5 mm. The corner radius of the
finished shell determines the punch radius on the second draw, which is 1.6 mm.

Fig. 10.38 Determining the relationship of punch-nose radius and die-radius

Determine Die Clearance:


Table 10.5 is used for determining the draw clearance (space between the punch and
die) of the shell in Fig. 10.33. Clearance for the first draw would be 1.09 x 3.2 = 3.488
mm. This will allow for thickening of the shell wall during drawing without galling. The
clearance for the second draw would be 1.1 x 3.2 = 3.52 mm, unless final sizing or
surface finish is important. When these factors must be considered, the clearance for
the second draw could be reduced to 1.05 x 3.2 mm = 3.36 mm or less, depending on
the ductility of the material.
Design of Sheet–Metal Bending, Forming and Drawing Dies 721

Determine Draw pressure:


The formula for estimating drawing pressure is
D
(
P = pdt (UTS) __ – C
d )
when,
Shell diameter, d = 50 mm
Material thickness, t = 0.8 mm
Ultimate tensile strength, UTS = 421 N/mm2 (from handbook or Table 10.6)
Blank diameter, D = 105 mm
Constant to cover friction, C = 0.7
Then,
105
( )
P = p × 50 × 0.8 × 421 × ____ – 0.7 = 74.066 KN ª 7.4 tons.
50
Blank-holder pressure may be figured as a maximum of one-third drawing pressure;
therefore, approximately three tons should be added to the drawing pressure for a
total of 11 tons. A 15-ton press would have plenty of capacity for the total possible
maximum pressure of 11 tons required for the operation.
At this point, it would be well to check the selection of material to see whether it will
withstand the drastic distortion imposed on it by the drawing operation. Table 10.2
shows the grades of low-carbon steel recommended for drawing round cups accord-
ing to cup reduction and punch-nose radius for specific sheet thickness. The punch-
nose radius for the first draw for the workpiece in Fig. 10.37 has been determined to
be 8.8 mm or 11t. According to Table 10.2, cold-rolled steel of commercial quality
would probably be satisfactory for a 40% draw. However, on the second draw when
the nose radius is 1.6 mm or 2t, the cold-rolled steel sheet must be of drawing quality
for a 20% draw. Therefore, to ensure the success of the total drawing operation, the
material purchased should be cold-rolled steel of drawing quality.

Summary

Blanking and piercing operations are often followed by other related sheet metal oper-
ations like bending, forming and drawing. This chapter deals with various sheet metal
shaping operations like bending, forming and drawing.
• Bending is defined as shaping of sheet metal around a straight axis which
extends completely across the material. The result is a plane surface at an
angle to the original plane of the workpiece material.
• Forming is similar to bending, except that the form or bend is along a curved
axis instead of a straight one.
• A drawing operation begins with a flat blank which is transformed into a cup of
shall.
• The two methods used extensively in presses are wiping die and V bending.
722 Tool Design

• During bending of sheet metal, elastic stresses remaining in the bend area
after bending pressure is released will cause a slight decrease in the bend
angle. Metal movement of this type is known as springback.
• Bend allowance is the length of the neutral axis between the bend lines, or
other words, the arc length of the bend.
• Different types of forming dies are solid form dies, pad-type form dies, curling
dies, embossing dies, bulging dies, coning dies, and assembly dies.
• Drawing operation is referred to as shallow drawing when the cup is no deeper
then half its diameter. Deep drawing is drawing of cup deeper then half its
diameter. Deep drawing is often accompanied by an operation called ironing,
which is defined as the reduction in wall thickness of a shell by forcing it though
a tight die.
• The diameter of the blank can be determined from formula depending on the
ratio of the shell diameter to the corner radius.
• Draw ratio is defined as ratio of draw height to shell diameter. Draw ratio is
important factor to estimate the number of draws required.

Questions
1. What are the three groups of metal-shaping operations?
2. Referring to Question 1, what is the difference between the groups?
3. When bending, why is the workpiece oriented so that the bend is made across
the grain produced in rolling mills?
4. What is the rule of thumb for the minimum bend radii for most annealed
metals?
5. When designing V-bending dies, what is the desirable width of the opening in
the V?
6. What is air bending?
7. What is the included angle of the punch and die for air bending mild steel?
8. What determines the radius of bends formed by air bending?
9. What is the major advantage of air-bend dies?
10. What type of press is generally used for air bending?
11. What are the pressure requirements for bottoming dies compared to air dies?
12. What is the advantage of bottoming dies compared to air dies?
13. How is the bend radius formed when using wiping dies?
14. What cause springback?
15. What is the simplest way of combating springback?
16. Generally speaking, what is the amount of springback for common materials?
17. What is the only sure and practical way to determine the necessary amount of
overbend?
18. What is corner setting?
19. How much should metal be sequeezed for corner setting?
20. What is ironing?
21. Why are U des (Channel dies) generally equipped with pressure pads?
22. Why is it difficult to develop general formulas for forming pressures?
Design of Sheet–Metal Bending, Forming and Drawing Dies 723

23. Why are forming dies generally mounted at one or more of the later stations in
progressive dies?
24. What problems may be encountered in the case of a misfeed when solid form
blocks are mounted in a progressive die?
25. Why should a radius be provided when the work material must move over a
form edge?
26. What methods are used to remove the workpiece form the punch or die cavity
after a forming operation?
27. What is the disadvantage when spring are used to provide pressure pads?
28. What is an air cushions?
29. What is the advantage of air cushions?
30. What are the advantage of hydropneumatic die cushions?
31. What is a curling die?
32. What is the purpose of curling?
33. Why should curling dies be polished and properly heat-treated?
34. What determines the size of curl?
35. What can be done with the metal tears or wrinkles with a particular design in
an embossing die?
36. What is the major difference between embossing and coining?
37. What are the major advantages of coining?
38. What type of press is generally used for heavy coining operations?
39. Why must coining dies be highly polished and free from scratches?
40. What is bulging operation?
41. Why is rubber or urethane preferred as a pressure-dispersing agent for bulging
operations?
42. Why must a fast-acting press be used when using a liquid as pressure-dispers-
ing agent for bulging operations?
43. Why is it necessary to make a provision to contain the metal for deeper
draws?
44. What is considered to be a shallow drawing operations?
45. Why is it necessary to anneal the workpiece between each drawn in an extreme
draw situation?
46. What is meant by ironing when discussing drawing operations?
47. Generally speaking, what force is required on the blankholder to prevent wrin-
kling during a drawing operation?
48. What will be the result if blank-holder pressure is too heavy?
49. Where do tears usually occur when blank-holder pressure is too high? Why?
50. What is the general rule for the size of the punch radius on draw dies?
51. Referring to Question 50, what must be done when the radius must be less
than four times the metal thickness?
52. What is the result when the radius on the draw ring is too sharp?
53. What is the general rule for the draw radius?
54. Why is it important not to stop once a drawing operation is started?
724 Tool Design

55. What is the one of major methods of reducing friction during a drawing
operation?
56. What type of lubricant is used when it is necessary to prevent galling and sei-
zure between the tool and the work material during a drawing operation?
57. How are pressworking lubricants applied?
58. What are the three principal classes of available low-carbon sheet steel for
drawing purposes?
59. Why is mill oxide (scale) on the surface of the work material detrimental to
deep draw?
60. What is the theoretical maximum per cent reduction that can be obtained in
one draw?
61. What is the maximum per cent reduction for one draw that can be obtained
under production conditions?
62. What is the drawing speed for low-carbon steel? For nonferrous materials?
63. What is the die clearance of a draw die?
64. Why is it generally necessary to leave clearance in excess of the metal
thickness?
65. What is a single-action draw die?
66. How is a flanged shell that cannot pass through the die removed from a draw
die?
67. How are deep draws made in single-action dies?
68. What is a double-action draw die?

Problems
1. Calculate the blank length of the workpiece shown in Fig. 10.39a.
2. What is the required bending pressure to air-bend the workpiece shown in
Fig. 10.39f. Die opening is eight times metal thickness.
3. Calculate the bending force for the workpiece shown in Fig. 10.39d. A U die will
be used to form the workpiece. Assume the die corner radius to be 3.2 mm.
4. Calculate the bending force required for the workpiece shown in Fig. 10.39b,
using a wipe-down die. Assume a wipe-down radius of 9.5 mm.
5. Determine the percentage reduction for the workpiece shown in Fig. 10.40a. Is
it possible to form the cup in one draw?
6. What is the depth (draw ratio) for the workpiece shown in Fig. 10.40a? Does
the depth of draw indicate that the cup can be formed in one draw?
7. What is the draw ratio of the workpiece shown in Fig. 10.40c? Can this cup be
formed in one draw?
8. What is the mount of force that will be required to draw the symmetrical cup
shown in Fig. 10.40a?
9. What will be the correct radius on the punch and die for the workpiece shown
in Fig. 10. 40a?
Design of Sheet–Metal Bending, Forming and Drawing Dies 725

Fig. 10.39 Bending problems


726 Tool Design

Fig. 10.40 Drawing-die problems

Design Problems

Use the design and drafting procedures given in Chap. 1 as a guide for the following
problems. A finished and complete tool drawing will be expected.
1. Design a combination piercing and bending die for the workpiece shown in
Fig. 10.39a.
2. Design a combination piercing and bending die for the workpiece shown in
Fig. 10.39b.
3. Design a combination piercing and bending die for the workpiece shown in
Fig. 10.39c.
Design of Sheet–Metal Bending, Forming and Drawing Dies 727

4. Design a bending die for the workpiece shown in Fig. 10.39d.


5. Design a combination piercing and bending die for the workpiece shown in
Fig. 10.39f.
6. Design a forming die for the workpiece shown in Fig. 10.39e.
7. Design a curling die to roll the raw edge of the workpiece shown in Fig. 10.40a.
Assume that edge has been lengthened enough to curl edge and hold present
overall dimensions.
8. design a combination piercing and forming die for the workpiece shown in
Fig. 10.41a. Use steel strip as work material.

Fig. 10.41 Forming problems


728 Tool Design

9. Design a forming die to form the workpiece shown in Fig. 10.41b. Use steel
strip as workpiece material.
10. Design a forming die to form the workpiece shown in Fig. 10.41c. Use alu-
minium strip as workpiece material.
11. Design a combination blanking, forming and embossing die for workpiece
shown in Fig. 10.41d. Use copper strip as workpiece material.
12. Design a draw die for one of the workpiece shown in Fig. 10.40 or 10.42.
Calculate the following in the early stages of the design procedure: (a) Per cent
reduction, (b) Depth of draw and draw ratio, (c) Number of draws required, (d)
Die clearance, (e) Blank size, (f) Drawing force, (g) Radius on punch and die,
and (h) Relationship of punch-nose radius and die radius.

Fig. 10.42 Drawing-die probelms

References
Books
• ASM: “Matals Handbook,” 4th ed., “Forming,” Metals Park, Ohio, 1967.
• ASTME: “Handbook of Fixture Design,” F W Wilson (ed.), McGraw-Hill, New york, 1962.
— “Tool Engineers Handbook,” 2d ed., F W Wilson (ed.), McGraw-Hill, New York, 1959.
• Eary, D F and E A Reed: “Techinques of Pressworking Sheet Metal,” Prentice-Hall,
• Englewood Cliffs, NJ, 1958.
• Paquin, J R: “Die Design Handbook, “The Industrial Press, New York, 1962.
• Schubert, P B (ed.): “Die methods, Design Fabrication, Maintenance, Applications,” Industrial
Press New York, book 1, 1960; book 2, 1967.
• Verson Allsteel Press Company: “Verson Die Manual,” Chicago, 1960.
• Wilson, F W: “Fundamentals of Tool Design,” Prenice-Hall, Englewood Cliffs, N J, 1962.
USING PLASTICS AS
TOOLING MATERIALS 11
11.1 INTRODUCTION
High production costs, close competition, and high investments have forced industry
to seek new materials and production methods in order to survive. Industrial periodi-
cals consistantly carry articles on cost-cutting methods and time-saving procedures.
Large industries throughout the country have completely reorganised or relocated
their plant facilities in order to reduce production costs or increase efficiency. Almost
any change in an industry’s pattern of operation, installation of new machinery, or the
adaption of new material can be pinpointed to one or two purposes to reduce costs
or to increase production.
One phase of production of primary interest to industry is in the initial tooling of the
product. Conventional industrial tooling methods are costly and time consuming.
Metals have been the main tooling material, and although their initial cost is not high,
the time spent in fabricating tooling from them is high and they are difficult to work.
Lead time for producing tools from metals to meet production requirements is high,
and in the event of an engineering or model change, the complete tool may have to
be scrapped.
Within the last decades, a new group of engineering materials and plastics have
come into existence. Some plastics have properties that make them acceptable as
tooling materials. Plastic tools are often equal to or better than metal tools in many
applications. In addition, many of the disadvantage of metals as tooling materials
have been overcome.
Plastics, as defined by the Society of Plastics Engineers, is “a large and varied group
of materials which consist of or contain as an essential ingredient a substance of high
molecular weigh which, while solid in the finished state, at some stage of its manufac-
ture is soft enough to be formed into various shapes most usually through the appli-
cation of heat and pressure.” This general property makes possible the formation of
various shapes without the machining operation generally required when metal tool-
ing materials are used. This results in the reduction in costs brought about by reduced
bad time, and reduced costs in a reduction in costs brought about by reduced dead
time, and reduced costs in reworking tools, low fabrication costs, ease of handling,
and reduced storage costs, to name a few.
Most tool designers have grown up with metal as the primary tooling material and
have developed a proficiency in this field that tends to close their minds to any other
method. The purpose of this chapter is to introduce the tool designer to the various
plastics used in tooling and the methods of applying them. The designer should
approach the subject with an open mind but should not allow himself to be carried
away. Plastics will not replace metal as a tooling material and certainly are not a
cure-all, but in many cases, they are applicable with considerable savings, especially
if they are viewed as an engineering material in their own light and not as a substitute
material.
730 Tool Design

11.2 PLASTIC COMMONLY USED AS


TOOLING MATERIALS
The major attraction of plastics is that they can be formed directly to a desired shape
by use of a master. Resins that form plastic tooling materials are therefore generally
purchased in a liquid or paste form. They can be worked at room temperature and are
made to take the shape of the workpiece by various techniques of laminating, pasting,
and casting. They are generally thermosetting and must be mixed with a curing agent
before they will set up.

Phenolics Phenolics, the earliest plastic tooling materials, were applied in the
aircraft industry as early as the 1930s. Liquid phenolics with acid catalysts cured
at room temperatures and were used for such tools stretch-form dies, large master
fixtures, and some foundry pattern work. Phenolics are very unstable and shrink
badly during curing and are therefore not suitable for quality tooling. In addition, they
are very corrosive to tools and checking instruments because of the acid catalyst.
Phenolics have largely been replaced by other plastics and are mentioned only for
their historical significance.

Polyesters The aircraft industry introduced polyester resins and reinforced them
with fiber-glass cloth. They are somewhat more dimensionally stable than pheno-
lics but still shrink in curing and do not maintain the dimensional accuracy required
for close-tolerance tooling. Originally used in large assembly and checking fixtures
where the light weight of a laminated construction was an advantage, they, too, have
been replaced by other plastics. Today, their use is limited to large moulds, such as
those for fibre-glass boats, when close tolerance is not important.

Epoxies Plastic tooling made great advances when epoxy resins were introduced.
They have almost completely replaced phenolic and polyester resins because of
their superior physical properties. They are available for both casting and laminating
application with practically no shrinkage. Since epoxy resins are the primary tooling
material in use today, the majority of the material in this chapter will be devoted to
their use.
The epoxies are thermosetting resins. In other words, after curing, they remain solids,
which may become softer when heated but will never again liquefy. In the uncured
state, they are either honey-coloured liquids or brittle amber solids which become
liquid when heated.
Enlarged 10 million times, the molecules of resin resemble short pieces of thread.
Once hardened (cured), these threads are joined together at the polymerised ends
and along the sides to form large cross-linked structure. Each molecule is tied to sev-
eral in a fishnet type of structure, which accounts for its great strength.
The curing (often referred to as hardening or polymerising) of epoxy resins is accom-
plished by adding an active reagent known as a curing agent (or hardener, or activa-
tor, or catalyst). Some curing agents promote curing by catalytic action, while others
participate directly in the reaction and are absorbed into the resin chain. Depending
upon the particular agent, curing occurs at room temperatures, with the heat produced
by exothermic reaction, or requires the application of external heat. Heat energy
always speeds the reaction. During the cure, two reaction are involved: conversion,
Using Plastics as Tooling Materials 731

or the actual disappearance of reaction elements, and cross-linking, or the coupling


of the threadlike molecules into three-dimensional networks through reactive residues
to form the desired thermoset resins. The physical properties of an epoxy casting
depend upon the extent of the cure.
The tool designer and the toolmaker are generally concerned with the practical
application of epoxy resins as tooling materials. Since they lack the time, funds,
or background for experienced work, they must rely on the manufacturer’s
recommendation for the type of resin and hardening agent for a particular application.
Many tool failures are due to improper material selection or improper formulation
methods. The manufacturers of epoxy resins have carried on a great deal of research
in the adaption of their product to specific applications. Detailed instructions can be
obtained from them for the kind of resin to use, the type of curing agent, the length of
cure time, the resin application, and the physical properties of the resins. In addition
to specific instructions for a particular resin, general instruction are given for the
fabricating procedure, type of available fillers, and proper type of reinforcements.

Urethane compounds The urethanes are a group of rubberlike materials frequently


referred to as elastoplastics. The are supplied in the raw state, as liquids for curing,
as putty for trowelling, or as solid bars that can be machined to shape. They can be
formulated to form a soft, highly flexible rubberlike material to a hard, semirigid solid
used for metal-forming dies.
Liquid and paste urethane is prepared for use like epoxies. The liquid urethane must
be mixed with a hardening agent before it will set, or harden. The mixture cures at
room temperature without heat or external pressure.
Urethanes are nonshrinking and have high tensile, tear, and abrasion resistance.
Paste urethanes adhere to a wide variety of materials, including the major metals.
The softer grades have considerable resiliency and flexibility. Once cured, they have
little or no tendency to flow under pressure, and once the pressure is removed, the
paste will resume its original shape. They also have good resistance to oils, greases,
and chemicals.
Softer grades of liquid or paste urethanes are often used for making flexible moulds
used for casting duplicate parts. The mould is made by pouring the urethane against
the original model, allowing it to cure, and then stripping it off. A release, or parting
compound must be applied to the model to facilitate removal.
The major application of urethane as a tooling material is in making metal-forming
dies because of its noncompressibility. When forces are applied to a urethane pad,
they are transmitted uniformly in all directions. By containing the pad, these forces
can be directed to form complex shapes without such undesirable effects as wrinkling,
puckering, misalignment or fracture. Soft metals are not marred or scratched when
formed by urethane dies because of their flexible nature.

11.3 APPLICATION OF EPOXY PLASTIC TOOLS


Epoxy tooling provides a simple method of reproducing compound curvatures. The
epoxy resins themselves are initially high priced compared to raw tool steels, but
the finished tools are almost invariably cheaper than their metal counterparts. They
require fewer man-hours to produce, and the high skill of a tool and diemaker is not
required. Lead time is reduced in the development of new models, and economical
732 Tool Design

prototypes are easily reproduced, so that various departments within an organisation


can work on new models simultaneously. Design changes can be quickly evaluated,
so that ideas and corrections of the design department can be fed back into the
original model.
Epoxy tools are repaired or altered with ease. The surface of the epoxy roughened
by sanding, then built up with an epoxy material from which the tool was originally
developed.
Epoxy do have limitations, especially in certain areas of metalworking. Matacoding
provides high compression strength, tensile strength, impact strength, abrasion
resistance, and elevated-temperature performance. The choice between epoxies and
metal tooling depends basically on the severity of the operation. Epoxy is generally
not a suitable tooling material when superior material characteristics are required.
Epoxies will creep under long and severe stress, although filling and reinforcing will
help control creep to some extent. They must also be used at a fraction of their ultimate
strength to prevent fatigue failures. They will wear rapidly under abrasive conditions,
and they are not strong enough to be employed as shearing tools.
Combining metals and epoxy materials often make excellent tools. For example, drop
–hammer dies have been made with cast-epoxy facings on Kirksite cores. Steel inserts
backed up by epoxy have made shearing tools. Metal rings have been embedded in
deep-contour press tools to prevent the tools from spreading. Metal inserts can be
cast into epoxy at greatest points of wear. With a little imagination, the tool designer
can come up with similar hybrids that produce excellent tools at a minimum cost.
Basically, the following types of tools can be produced by use of epoxy resins.

Stretch dies Stretch dies are male dies over which sheet metal is formed by gripping
it on its edges and pulling or stretching until the shape of the die is reproduced in the
metal sheet. They can be made by laminating or casting, depending upon the size of
the die and load requirements. The dies are generally backed up by a core to reduce
material costs.

Draw and form dies Plastic draw and form dies are used in the same manner as
conventional draw and form dies. They are run in double- or triple-action presses
using conventional draw-die techniques. This type of tooling is subjected to high
stress, wear, abrasion, or heat. Filler materials are often added to the epoxy to enable
it to withstand the heavy service conditions. Steel draw rings may be applied at points
of high stress and abrasion.

Hydraulic-press dies Hydraulic-press forming is technique in which a thick rubber


pad is substituted for one component of the die. A male punch or a shallow female die
is held upon a fixed base, and sheet metal is shaped against it by hydraulic pressure
transmitted through the rubber pad. Dies are commonly made by epoxy-casting
methods, the die being backed up with steel or hardboard baseplate. Large dies may
be cast with a core to conserve the epoxy material.

Drop-hammer dies Dies of this type are used to cold-forge sheet metal into shape
by using a drop hammer. The dies are made in two parts a female lower part contain-
ing the die cavity and a male upper part containing the matching punch. The punch
is dropped to ram the sheet metal into the cavity. Since this type of die is subjected
to shock loads, it is generally constructed of a low-temperature alloy core, such as
Using Plastics as Tooling Materials 733

Kirksite, faced with a plastic. The female die is faced with a hard-surface epoxy, while
the male is faced with a more resilient epoxy. The resilient face has a tendency to
deform and aid in the even distribution of the impact force.

Checking fixtures Checking fixtures are master forms used to check the shape
of tools and assembled parts produced in these tools. The primary requirements
for plastics used for this application are ease and accuracy of duplicating and
dimensional stability. Laminated epoxy resins using fibre-glass cloth, rovings, and
mat are construction types used for checking fixtures.

Mockups and moulds Mockups and moulds are often made of a proposed product
for improval of appearance or for engineering study. For the most part, they are
nonfunctional. They are generally constructed by using templates spaced at specific
locations to define the required contour. The templates are mounted on a metal
baseplate. Large mockups require a steel weldment base. Holes are drilled near the
edge of the templates, and wires are strung through the holes to define the shape of
the mockup between the templates. A flexible wire screen is placed over the wires to
which the plastic material is laminated.

Assembly fixtures An assembly fixture is a structure that finished workpieces in


correct relation to each other for final assembly. Plastic assembly fixtures are generally
a framework constructed with epoxy lamination resin, glass cloth, and plastic tubing.
The advantage of plastic over metal is its light weight and the ease of constructing
compound curved surfaces. Metal clamping devices can easily be incorporated in the
lay-up.
Plastics are often used to form compound curved surfaces within a metal assembly
fixture. The basic structure is metal, and contoured locating pads are made of plastic
cast against a master model.

Drill jigs Drill jigs are made by potting a drill bushing directly into a laminated
structure or by installing the bushing directly in the fixture at the time of lamination.
In the potting technique, an oversize hole is drilled into the supporting structure, the
bushing is located in by reference to a master, and epoxy resin is poured into the hole
and allowed to cure. In the laminating process, dowel pins are used to locate the drill
bushings on the surface of the master. The bushings are then installed permanently
during the lay-up procedure.

Machining fixtures Epoxy resins may be used to good advantage when constructing
the nest in conventional milling fixtures, lathe fixtures, boring fixtures, etc. This
practise is especially helpful when the workpiece to be machined contains irregular or
compound locating surfaces. The metal fixture is constructed by conventional fixture-
fabrication methods with a pocket or open area left for the epoxy resin. Paste epoxy
is trowled into the area, and the workpiece is pressed into the epoxy to form the nest.
The workpiece must be coated with a parting compound to prevent sticking. The
epoxy is allowed to cure with the workpiece in place. The workpiece is then removed,
leaving a nest the shape of the workpiece.

Duplicating models Epoxy plastic are excellent for constructing master models
for duplicating mills and pantographs. Exact copies can be taken from masters, and
epoxies far surpass wood or plaster for dimensional accuracy and durability.
734 Tool Design

11.4 CONSTRUCTION METHODS OF


PLASTIC TOOLING
One of the major advantages of plastics as tooling materials is the ease in which they
can be shaped. This is due to the fact that they are liquid in the raw state at room
temperature. Most methods of construction are therefore are some form of a casting
process. These methods are generally broken down into four basic techniques,
laminating, surface casting, mass casting, and paste construction. At times, it is
necessary to use two or more of these methods in the same job.
The construction technique depends upon the size, type, and application of the tool.
Table 11.1 will be helpful in determining which method to use but is intended only
as a guide. Specific conditions in any particular application or plant may alter these
factors.

Table 11.1 Desirability Factors* of Plastic Tooling Methods


Criterion Laminate Metal core surface cast Mass cast Paste

Dimensional stability 1 3 4 2
Shrinkage (during cure) 1 3 4 2
Least weight 1 4 3 2
Labour cost 4 3 1 2
Material cost per lb 4 3 1 2
Strength 1 2 3 4
Toughness 1 2 4 3
*The table is intended to be used only as a guide. Specific conditions in any particular application or plant may alter
these factors. Desirability ratings: 1. most desirable; 2. satisfactory; 3. fair; 4. least
SOURCE: Ren Plastics, Inc.

Mould or model preparation Plastic casts or laminates can be taken from wood,
plaster, plastic, or metal surfaces, provided the surface is properly prepared. The
preparations described in the following paragraphs generally apply to the preparation
of a master surface to be used with any of the four basic construction techniques.
Wood masters should be finished with a lacquer sealer. The surface of the model
should be cleaned thoroughly, using acetone, alcohol, or other commercial solvent.
The surface is then coated with a high-grade carnuba-base paste wax and thoroughly
rubbed out with a lint-free cloth.
The next step is to brush or spray on a commercial parting agent (generally a vinyl
alcohol). Spray coats should be allowed to dry 20 min between coats; brush coats will
dry in 1 or 2 hr. The number of coats required will depend upon the condition of the
master. After the parting agent has dried, another coat of paste wax should be applied
and rubbed out with a lint-free cloth. The model is now ready for the casting operation
(see Fig. 11.1a).
Master made of the commonly used plasters are similarly prepared, with one or
two slight variations. The plaster master is first well sealed with a lacquer sealer
and allowed to dry. Wax is applied and rubbed out thoroughly with a lint-free cloth.
Commercial parting is brushed or sprayed on. After the parting agent has dried, paste
wax is again applied and the excess is wiped off. The wax should not be rubbed out
(see Fig. 11.1b)
Using Plastics as Tooling Materials 735

Fig. 11.1 Preparation of the master mould or model (Ren Plastics, Inc.)

Plastic masters are prepared by first removing all foreign matter from the surface of
the model and applying a coat of wax. The wax is then rubbed out and a contrast-
colour commercial parting is applied. After the parting has dried, another coat of wax
is applied and the excess is wiped off.

Laminating A laminated plastic tool is made of alternate layers of glass cloth


and liquid laminating plastic. After the laminations are completed, the liquid plastic
solidifies into a strong, rigid form. The finished piece has the exact size and shape of
the surface from which it was moulded.
Laminated tools are usually reinforced with, or become a part of, a framework. This
framework can be made of plastic material, fabricated steel, or aluminium, and the
type of framework developed depends upon the end use of the tool (see Fig. 11.2).

Fig. 11.2 A die-model duplication of automotive floor pan used for kellering operations.
Contains laminated face with glass-tubing framework. (Ren Plastics, Inc.)

Laminated plastic tools are used wherever accuracy, strength, and weight are prime
considerations. They possess the best dimensional stability of any plastic tools and
are widely used in industry for gauge tools, checking fixtures, and inspection fixtures.
Laminated plastic drill fixtures cannot be surpassed for accuracy, cost, or time to build.
They are light weight and durable and will not rust or warp with age.
Successful tooling procedures with laminating plastic materials are not difficult if basic
rules are followed and if accurately measured proportions of quality materials are
used.
736 Tool Design

The laminating methods consists basically of building up alternate layers of glass


cloth and plastic to the form until the desired thickness is obtained. Each layer of
0.3 mm thick glass cloth with the laminating material lays up to approximately 0.5 mm
thick. Hence, 18 layers equals 9 mm.
The following procedure is a guide for lay-up of laminated plastic tooling:1
1. Apply the parting agent to the surface of the
model according to instructions for mould
and model preparation.
2. Build-ups for parting planes, run-out aprons,
or framework may be required. Make them
of clay, plaster or wood (see Fig. 11.3).
3. Apply parting agent to surface of build-ups.
4. Apply desired surface-coat plastic according
to instructions on can. Fig. 11.3 Build-ups used to con-
fine the plastic lay-up
5. Wait until surface coat has become tack free.
to a certain area (Ren
(If material can be dented but does not stick
Plastics, Inc.)
to the fingertip, it is considered tack free.)
6. Using a glass-paste mixture, fill all sharp corners and detail which might cause
voids or bubbles under the first layer of glass cloth. (Glass paste is made up of
laminating plastic chopped glass fibre or floc. In mixing, add sufficient dry fibre
and cotton floc to obtain a smooth paste. Strength is obtained from the fibre,
while cotton floc produces a smooth, viscous mixture. Adjust quantities of each
to suit the job. Do not use heavy sections of the paste, as the shrinkage of this
material is greater than a laminated plastic.)
7. Apply alternate layers of laminating resin and glass cloth to the model until
desired thickness is completed.
8. Small patches of glass cloth applied first (near the surface coat) will help pre-
vent voids where the cloth tends to bridge over inside radii or sharp inside cor-
ners. Never allow succeeding patches of cloth to have their joining lines directly
over the joint lines of the previous layer. This could weaken the laminate and
possibly set up a flaw line through the tool.
9. Try to make the plastic work up through the glass cloth as you proceed by
applying dry cloth and dabbing or stippling with the paint brush. After the plastic
has soaked up through the cloth, additional plastic is applied with the brush for
the succeeding layer.
10. The last few layers of glass cloth are usually larger pieces, cut to cover the
entire surface of the tool if practical.
11. Do not apply more than 18 layers (9.5 mm thick) at any one time since the heat
generated by the plastic can cause excessive shrinkage or warpage of the
finished tool. If excessive heat is noted at any time, it is a good rule to stop until
the material tool. If excessive heat is noted at any time, it a good rule to stop
until the material sets up and has cooled off. If the surface is glossy, it should
be sanded before the next layer is applied.

1
Ren Plastics, Inc,. Lansing, Mich.
Using Plastics as Tooling Materials 737

12. If a framework is to be added to the laminated tool facing, this should be done
immediately after laminating, before removing the plastic tool from the model.
As a general rule, wait at least 6 hr after the last liquid plastic has been applied
to the tool before removing it from its model.
A laminated tool is usually strengthened by adding a plastic framework attached to
the laminated face of the tool. The framework generally must be fastened to the tool
before it is removed from the model or mould to prevent distortion, especially with
large flexible laminates. The amount and type of supporting structure depend upon
the size of the tool, the function of the tool, and the amount of rigidity required.
The framework is best made of material exactly like the tool facing. Such factors as
temperature change, moisture, age, and weight affect plastic tools having frameworks
made of materials which are dissimilar to the laminated-tool having framework made
of materials which are dissimilar to the laminated-tool facing.
Square and round plastic tubes make an excellent framework when used as recom-
mended. This tubing material can be cut to fit; when it is applied to the laminate with
glass cloth and plastic, it becomes an integral part of the tool. Round tubes can be
used to fit a curved structure by spiral cutting, as shown in Fig. 11.4. Other tubing
applications are shown in Fig. 11.5.

Fig. 11.4 Spiral cutting of round plastic tubing: (a) Angle of saw table determines spacing
of cuts. Tilt saw table approximately 25o and secure blocking material in back
of saw blade with sufficient distance so that saw blade will not cut more than
1.6 mm more than round tube-wall thickness (b) Angle of guides must be sufficient
to allow tube to clear saw frame (c) Finished cut tube should appear as shown. It
can be easily bent to conform to quite severe contours. Position the tube on tool
and secure with RP-1135 Quick Set. Apply two or three layers of glass-cloth tape
and laminating mix over the tubing and onto the attached surface (Ren Plastics,
Inc.)
738 Tool Design

Fig. 11.5 Typical plastic-tubing applications: (a) Reinforcing the flanges of the skin with
spiral-cut tubing (b) Bonding base structure to skin using round Ren tubing
(c) Bonding the reinforcing ribs to skin (d) Typical tubing joint wrapped with glass
tape. RP-1710 is applied with brush as tape is wrapped. Laminate usually picks
up around the edges slightly when removed from the master and will not rematch
unless proper webbing or tubing reinforcement is added to hold it in the exact
contour position (e) Bonding a spiral-cut round tubing to skin for reinforcement
(f) Another application of spiral-cut round tubing (Ren Plastics, Inc.)

Figure 11.6 shows the method of constructing a joint made with square and round
plastic tubes. The surface of the tube is sanded when bonding is disired. The tubes
are cut to a reasonable fit and bonded into place with a plastic paste. The joint is then
wrapped with glass-cloth ribbon preimpregnated with laminating plastic material.
Other types of frameworks, can be made from sheet materials such as glass-laminate
honeycomb structures, paper-base phenolic sheet stock, plywood, or welded steel.
Using Plastics as Tooling Materials 739

Fig. 11.6 Constructing a joint with Fig. 11.7 A surface-cast aircraft leading-edge
square and round tubes stretch block. The core is Kirksite
(Ren Plastics, Inc.) with epoxy surface-cast. (Ren Plas-
tics, Inc.)

Because the choice of framework has a great effect upon the dimensional stability of
the tool, it is especially important that appropriate consideration be given to this phase
of the tool-building procedure. Movement within the framework can cause warpage in
the laminated-tool facing.

Surface casting A surface-cast tool usually consists of a metallic core, rough-cast


to the general shape of the finished tool. This core is suspended over a model of the
working face of the tool, and liquid plastic is then cast into the space between the
model and the metallic core. The core may be an aluminium or Kirksite casting, a
casting of sand and gravel with epoxy binder, or even a rough-out hardwood plug.
A number of epoxy formulations are available for this use, depending upon the tool
requirements.
Tools made by the surface-cast method are used in almost every industry involved
with metal forming. A wide range of physical properties is available in commercial
cast plastics. For drop-hammer dies, both hard and resilient materials are available.
A resilient epoxy will prevent chipping of small radii in this type of operation. For draw
dies, a plastic which will allow good metal flow is available. Forming dies and bulging
dies require a tough wear-resistant surface, and here again plastic surface-casting
material is available. In any application where metal-forming tools are required and
the production run is limited, surface-cast plastic tooling should be given serious
consideration. Fig. 11.7 shows a typical surface-cast tool.
The surface-casting method is also widely accepted in the foundry patternmaking
trade. The savings gained from using the surface-casting technique are found in
the finishing and hand-barbering operations usually associated with metal patterns.
Plastic patterns are extremely useful for duplicating since they are cast to the exact
size and finish of the mould from which they are made.
740 Tool Design

Surface-casting plastics can be cast in thicknesses varying from 3 to 20 mm without


generating excessive heat or shrinkage. Casting thickness depends upon the size of
the tool and the method used in surface casting.
Three different methods have proved successful in surface-casting work.
1. Pour method The core is suspended over the die model and sealed around
the edges. The plastic is poured through spouts or at an open end. The core
is usually elevated at one end. Figure 11.8 shows the basic steps in the pour
method.

Fig. 11.8 Basic steps in pour method of surface casting: (a) Molten metal cast into sand
mould to make the core (b) Plaster mould being made on model of wing leading
edge (c) Metal core suspended for set-up prior to plastic cast (d) Pouring plastic
into space between plaster mould and metal core. Ends sealed to retain plastic
(e) Finished tool ready to be mounted on press for stretch-forming metal (Ren
Plastics, Inc.)
Using Plastics as Tooling Materials 741

2. Squash method Liquid or paste plastic is placed on the die model. The core is
then placed on the material and allowed to settle to a predetermined position.
3. Pressure-pot method The core is suspended over the die model and sealed
around the edges. Liquid casting is forced into place at a low point and allowed
to vent at high points.
The procedure for the pour method is described in the following steps. The core is
prepared in the same manner for all surface-casting methods.
1. The core must first be made and this is usually accomplished in the Kirksite
or aluminium foundry. A rough pattern is made to cast the metallic core. The
working face, i.e., the face on which the plastic surface is to be cast, is made
approximately 12.5 mm smaller than the finished surface (to allow space for
the surface-casting plastic). This pattern is then rammed into foundry sand,
and the Kirksite or aluminium core is cast. If a plastic core is required rather
than Kirksite or aluminium, refer to the mass-casting subsection, below.
2. The aluminium or Kirksite core can be used as soon as it has cooled to room
temperature. The working face is usually cleaned by sand blasting. If the core
is not to be used within 24 hr after sand blasting, paint the surface with a
liquid tooling plastic (laminating or casting type) to prevent surface oxidation.
When ready to use, the painted surface should be sanded and washed with
denatured alcohol.
3. Prepare the surface of the die model or plaster splash in the same manner as
for laminating (see the subsection on mould and model preparation).
4. Depending upon the surface-cast method to be used, set up the core and die
model for the casting operation. It is usually advisable to try the core in the
mould before the resin is added, especially when the squash method is being
considered. This should be done before mixing the plastic materials, i.e., a dry
run.
5. Check set-up to see whether air could be entrapped by the casting plastic as
it enters the mould. Drill vent holes through the core in these areas or elevate
one end of the set-up to eliminate the possibility of any air entrapment. Always
have the core on top of the die model, since this will keep any small air bubbles
away from the working face of the tool.
6. Mix plastic materials thoroughly according to direction on the can. When pouting
into the mould, allow a steady stream of material to flow until the complete cast
is finished. Do not change position of pouring, as this may entrap air.
7. After proper curing period, remove tool from die model and complete.
The pressure-pot system may be considered as an alternate method for surface cast-
ing utilising pressure instead of gravity force for proper material flow. Care should
be exercised to make sure that everything is in readiness before mixing resin and
hardener or before placing entire mixture in pressure pot. The selection of the proper
material is important in this procedure, just as it is in other applications. Prime factors
to be considered in advance include thickness, mass, material to be cast against, and
the desired results. It is recommended that the pressure pot be placed as close as
feasible to the inlet of the tool to eliminate excessive hose cost. Pot and hose should
be thoroughly cleaned as soon as the entire mass of material has been pumped into
the tool. Clean the system by allowing a circulation of isopropel alcohol throughout the
system until material is well mixed with the alcohol. Drain off and finish-clean by hand
(see Fig. 11.9 for pressure-pot setup).
742 Tool Design

Fig. 11.9 Set-up for pressure-pot system of surface casting (Ren Plastics, Inc.)

Mass casting At first glance, the casting of a plastic tool in a solid mass seems to
offer the greatest saving in time and material. However, closer examination often
indicates that this is not the best method.
Because the preparation of the mould or model is the same for mass casting as it is
for laminating, it appears possible to save the long and costly labour hours involved
in laminating. In addition, mass casting also eliminates the cost of producing and
preparing the metal cores needed when the surface-casting method is utilised. Unless
these savings do occur, there is no there is no reason to build a mass-cast tool, since
this type of tool has some disadvantages.
Probably the greatest of these disadvantage is the low-strength properties of most
mass-casting materials. Even the addition of fillers such as sand and gravel, glass
balls, volcanic ash, pumice, cork, and ground nutshells do not appreciably increase
the physical properties of this type of material.
Another possible disadvantage is the high shrinkage encountered, particularly in the
heavier casting, due to the heat generated when the casting is cured.
In the past, this shrinkage has made it advisable to build a laminated surface first
and then back it up with a casting material or cast the tool oversize and rework the
surface. Recent developments in epoxy casting plastics have greatly reduced both
these objections. These newer developments offer a considerable increase in physical
properties while permitting castings up to 300 mm thick without generating excessive
heat with the accompaning high shrinkage.
The preparation of the mould for mass casting is the same as for surface casting.
Following this, the required amount of material is poured into the cavity, allowed to
cure for the necessary length of time, and then from the mould (see Fig. 11.10). The
finished tool is shown in Fig. 11.11.
Using Plastics as Tooling Materials 743

Fig. 11.10 Typical set-up for mass-cast metal-forming tool (a laminated surface also may
be used on these tools). (a) Die model. (b) Build-ups of wood added to establish
parting plane for die halves. (c) Retaining box to establish outside dimensions
of die. Corner fillets, thickness wax, and positioning keys may be added at this
point if required. Prepare surface with parting agent for plastic cast. (d) Pour-
ing Ren mass-casting plastic. (e) Die half completed. Reference line scribed on
sides of die. Jackscrews adjusted to establish plane of base. (f) Die half inverted
on surface plate. Surface cast to die half using squash method. Jackscrews settle
down through soft plastic to touch surface plate. (Ren Plastics, Inc.)

Fig. 11.11 Mass-cast aircraft stretch block. Block is cast solid with epoxy resin mixed with
sand gravel. (Ren Plastics, Inc.)

When an unusually large mass is being poured (over 150 mm thick), it is often desirable
to add one of the fillers mentioned above. This prevents a high exotherm and greatly
reduces the amount of shrinkage during cure. It also results in a lower cost per cubic
foot of cured material. Wood flour incorporated into the filler will produce an excellent
surface finish.
The same method is used when casting a plastic core for a surface-cast tool, except
that it is necessary to cast the core approximately 12 mm smaller than the desired
finished surface to allow space for the surface-casting material.
744 Tool Design

The mass casting of heat-cured, high-temperature materials makes possible the use
of plastics in applications formerly considered too severe for plastic tooling.
The epoxy-alloy method of producing high-production plastic dies involves the use of
epoxy resin filled with chopped fibre glass and chopped steel fibre, cured under heat
and pressure. While dies produced by this method are more expensive than ordinary
plastic dies, they have shown some remarkable results in production runs.

Paste resins Paste resins for producing plastic tools are useful in almost every phase
of the metalworking industry. The principal advantage of a paste resin is that it is a
nonflowing material. It can be applied to vertical surfaces of a die model in the same
manner as a laminate. It does not have to be cast to a level surface like a liquid, nor
does it have to be retained in position with special equipment. It is perfectly suited for
the squash method of surface casting. It can be trowelled into place or applied under
pressure as a caulking operation.
The aircraft industry uses paste resins in making spline master models, which consist
of a series of templates set upon a common base plane. The top edges of the tem-
plates define a three-dimensional shape of some aerodynamic portion of an airplane.
It is necessary to fill the spaces between these templates with a dimensional stable
material which can be shaped, or splined, to a very smooth surface and will then
solidify without movement, to produce an accurate three-dimensional surface.
The principal difference in the use of paste and liquid resins lies in the mixing. The
resin and hardener portions of these materials are combined on a flat surface using
wide-blade putty knives. The resin and hardener are different colours so that complete
mixing is assured. Figure 11.12 illustrates the mixing and use of paste resins.

Construction by trowelling and splining This type of construction is used primarily


when no master model or pattern is available. In other words, it is used to make
the master. Examples include parototypes, master models, Keller models, casting
moulds, foundry patterns, and showpiece models. Basically construction by trowelling
and splining consists of building a skeletal structure from templates spaced at specific
locations to define the required contour. The templates conform to rigid dimensions
at known intervals. The area between the templates is then filled with plastic to form
a smooth-contoured surface:
Considerable skill and experience are needed when building this type of tooling. The
following suggestions1 should serve to familiarise the tool designer with the basic
construction techniques.
1. The first assumption in constructing a spline master model is that accuracy is
of prime importance. The base structure, being the foundation of the model,
must be accurate, stable, and self-supporting, especially if it is to be moved.
Stress-relieved, welded-steel construction is common for this purpose, as are
plastic-faced granite slabs.
2. The template material should be dead-soft, stretcher-leveled steel sheets, filed
and checked to accurate, loft lines. An extra set of 9.5 mm holes punched or
drilled in the templates (before filing) is required to hold the plastic surface (see
Fig. 11.13a).

1
Ren Plastics, Inc., Lancing, Mich.
Using Plastics as Tooling Materials 745

Fig. 11.12 Mixing and using paste resins: (a) Measure proper proportions of resin and
hardener into two piles on flat surface. (Note the use of two putty knives.) (b)
Mix resin and hardener thoroughly. (c) Apply the mixed paste as a filler material
to fill a sharp radii. (d) Optional use of a caulking gun to apply mixed paste.
(e) Using paste material in a squash method to make impression of any definite
form [see (f)]. (f) The final impression with form removed after past has set up.
(Ren Plastics, Inc.)

3. This operation is not accomplished in the same manner in all aircraft plants,
but the end result is similar. The templates are held in a common plane (usu-
ally perpendicular to the base) and on a reference line established on the base
line. The complete series of templates makes up the desired three-dimensional
surface (see Fig. 11.13b)
746 Tool Design

Fig. 11.13 Construction by trowelling and splining (Ren Plastics, Inc.)

4. Once the template set-up is completed and inspected for accuracy, the plastic-
tool builder takes over. Strips of 3 mm square-mesh galvanised hardware cloth
are cut to fit the spaces between the templates the templates. The hardware
cloth is laid in the spaces and pressed down below the rows of 9.5 mm holes.
Next 6 mm steel rods (clean roughened surfaces) are slid into the 9.5 mm
holes. Roads can run from one end of the set-up to the other if holes are in
line. Otherwise, they must be at least the length of two template spaces. The
rods must lie loosely in the templates and must not cause the template edge
Using Plastics as Tooling Materials 747

to move out of its normal position. Next the wire cloth is pulled up behind the
rods and attached to the rods with hog rings or wire ties. One tie in every other
spaces and on every other rod is sufficient. (see Fig. 11.13c).
5. Two separate and distinct paste plastics are required to fill the surface of the
model. The first is nonflowing paste, which is used on the model first to fill the
surface roughly. This material may be termed the base caot. The second mate-
rial is a more flowing paste which is used only as a finish coat on top of a base
coat and its thickness should not exceed 1.6 mm. One thing must be remem-
bered throughout the application of the plastic on the master model: uniform
thickness of the base coat and finish coat are important. A variation of 3 mm.
total thickness can cause excessive heat and warpage of the plastic surface.
6. The base coat is applied after mixing (see instruction on containers) with a
putty knife or spatula. The first application is intended to cover the rods and
wire cloth without pressing a large amount of material through the cloth. Merely
try to fill all openings in the surface. Proceed until the complete area is covered.
A slab of hard rubber (70 to 80 durometre) 12 mm thick by 75 by 250 mm long
is helpful at this point in cleaning the template edges and preventing excess
material from projecting above the finish surface.
7. When the first coat becomes hard (usually the area covered first is ready
immediately), a second layer of base coat is applied. This time fill each space
up to the template edges.
NOTE: You will find from this point on that it will help not to add plastic to
adjacent template spaces but to work in every other space. This allows
the material in one space to be worked without disturbing the material in a
finished section.
8. Use the rubber slab to remove the excess paste plastic to the model. The
rubber slab will remove enough excess plastic so that the cured surface of this
second application will be approximately 0.8 mm below the template edges.
Fill the entire model surface in this manner and allow the material to harden
(about 1 hr).
9. The finish application, using the finish coat, is applied in the same general
manner as before, working every second space to allow the material to cure
before disturbing the finish surface. The spring-steel spline tools are now used
instead of the rubber slab for removing the excess material. When applying the
white finish plastic, always “butter” the entire surface to be worked with paste
plastic. This operation is very much like icing the cake. After the paste plastic
is applied, the excess is removed with the spline tool. Do not try to achieve a
perfect finished surface on the first application. Try to fill all the areas up to the
template edges without having any material projecting above. Once this opera-
tion is complete, the model should set overnight or 12 hr, to allow a complete
cure out of the base materials. The splining technique is shown in Fig. 11.14.
10. The final work on the surface may take as much as three applications of the
finish coat, depending upon the skill and experience of the tool builder and
the accuracy required on the surface. Each application may only be a few
thousandths of an inch thick and the surface quality will improve with each
application.
REMEMBER: Apply material to the entire area before splining and always
remove all excess material with a spline tool. Avoid scraping the surface
after the plastic has cured.
748 Tool Design

Fig. 11.14 Use of splining tachnique on aicraft spline master model (Ren Plastics, Inc.)

11. Keep the splining tools clean by wipping with denatured alchol. Clean the mix-
ing surface before. Weigh out material to avoid any hard particles in the mixid
plastic. Weigh out materials carefully and use two putty knives for greater ease
in the mixing opreation. (One knife will clean the plastic paste off the other,
etc.) The materials are colour – coded to assure complet mixing. Mix until there
are no streaks of colour in the mixed plastic.
12. Always control the shape of the spline tool by touching three or more adjacent
templates when sweeping over a surface. Hold the tool at an angle to the sur-
face of about 45°. The spline tools should have rounded edges for this work
instead of two sharp corners, as shown in Fig. 11.15.

Fig. 11.15 Spline tools and their use: (a) Press down with fingers at each template while
sweeping surface. (b) Spline tools, various sizes and thickness required. Tools
made of strips steel. (Ren Plastics, Inc.)

Construction of drill jigs Drill jigs are constructed by laminating or casting drill
bushings in place during the casting or laminating process. Special bushing with a
knurled, grooved, or serrated body are used with this type of tooling (see Chap. 7).
Figures 11.16 and 11.17 show the basic methods of constructing drill jigs by casting
or laminating methods decribed are basic in nature and are intended only as a gen-
eral guide. Many variations are used, depending upon the material chosen and the
requirements of the intended application.
Using Plastics as Tooling Materials 749

Fig. 11.16 Construction of drill laminating: (a) Prepare form; a suitable form may be a
sample part or metal. Locate and install precision locating wood, plaster, plastic,
or metal. Locate and install precision locating pins of exact hole size required.
Thoroughly coat pins and surface of form with a recommended parting agent. (b)
Precut glass cloth to size. Holes may be cut or punched to fit locating pins. (c)
Using brush or spatula, apply thin layer of special surface-coat resin. Smooth out
all air pocket and voids. (d) When surface is tacky, brush on thin coat of laminat-
ing resin and apply first layer of glass cloth. Smooth out wrinkly,taking care not to
stretch cloth. Continue applications of resin and cloth, rotating weave of cloth as
each layer is applied, until desired thickness is obtained. (e) After curing, remove
laminated jig from and bore holes for drill bushing. Holes for drill bushings. Holes
should be at least 1.6 mm larger than OD of bushings. If chip clearance is not
desired, bore holes deep enough to seat bushing but not completly through and
washers (coated with a parting agent) placed under bushings of correct ID on
locating pins; selct bushings specifically designed for plastic tooling. Fill cavity
between bushing and jig with potting compound. Do not cover tops of bushings.
(g) Finished jig curing of potting compound. (Amecrican Drill Bushing Co.)

11.5 METAL-FORMING OPERATIONS WITH


URETHANE DIES
Urethane elastomers are used for metal- forming operation in a manner similar to that
of rubber metal-formating dies, and they offer similar advantages. The workpiece is
not scratched, die shapes are easily interchanged, alignment and mismatch problems
are eliminated, and tooling can be econmically constructed. However, urethane plas-
tics exceed die rubber in toughness, durability, and stress capacity. These characteris-
tics permit the precision forming of parts of higher definition with reduced springback
and distortion.

The urethane-die mechanism In contrast to convention metal-forming urthane tool-


ing employs only a single rigid punch. The female die is generally a simple, rectangu-
lar block of urethane that requires no specific shaping. When the unit is closed, the
750 Tool Design

die begins to deform at the start of the stroke and gradually assumes the shape of
the punch, providing the female die is contained. Since urethane is noncompressible
(the defliction occurs. The urethane die wrap the workpiece blank like a “solid liqued
“when that springback is virtually eliminated in many applications. Upon reaction of
the punch, the urethane die wraps the workpiece blank around the punch so tight that
springback is virtually eliminated in many applications. Upon reaction of the punch,
the urethane die promptly returns to its original dimensions.

Fig. 11.17 Construction of drill jigs by casting: (a) Thoroughly coat surfaces of sample part
or model with suitable parting agent. (b) Insert precision locating pins of correct
ID on locating pins; select bushings designed specifically for plastic tooling.
Washers (coated with a parting agent) may be placed under bushings to provide
proper chip clearance. (d) Fabricate a suitable form. Baseplate and strap may be
steel may be steel or other suitable material. Coat baseplate with parting agent.
Place assembled parts inside form. Shim face of part that contacts strap with
a piece of thin vellum or cellophane for clearance. slowly pour liquid plastic
material, avoiding air pockets and void; fill to top of strap. Do not cover top of
bushings. (e) Finshed drill jig after curing. (Amercan Drill Bushing Company)

General design considerations for urethane dies Since the urethane die (pad)
flows evenly away from the force displacing it, it is necessary to direct the majority of
the resulting forces against the forming areas of the punch. This is accomplished by
rigidly constraining the die in certain areas and relieving it in others. Metal die retainers
and deflector bars are used as constrictors. Relief is accomplished by providing an air
channel beneath the die pad to help constrictors. Relief is accomplished by providing
an air channel beneath the die pad to help control deflection and permit deeper
penetrations. Specific applications of constriction and relif are shown in the various
illustrations of urethane-die designs.
Using Plastics as Tooling Materials 751

Replace or interchangeable solid urethane die pads are generally best suited for
forming applications. Solid-urethane pads are obtainable commercially in various
sizes and thicknesses and are easily cut to the needed diamensions. Since there are
several families of urethane elastomers and not all are suitable for metal forming, the
beginner should depend upon the manufacture’s recommendations. Since there are
several famillies of urethane elastomers and not all are suitable for metal forming,
the beginner should depend upon the manufacturer’s recommendation for specific
applications. Die pads are selected in accordance with the necessary side and
bottom pressures to form a particular range of metals, metal gauges, and shapes.
Manufacturer’s representatives are well informed and equipped to make judgments
on this basis.
Sharp punch ends will eventually penetrate urethane and start a cut that result in
pressure loss. For long pad life, sharp ends should be stoned off to minimise this
problem. A loose urethane wear pad applied on top of the retained die pad will also
protect against cutting.
For longer life, urethane die pads should not be deflected beyond the maximum
deflection limit for that particular grade. The service life will corresponded to the
amount of strees (or overstress) put into the pad. For example, U forms cause less
wear; more complicated forms and short flanges cause greater die-pad failure.
Be sure that metal die retainers are strong enough to withstand the high side pressures
transmitted during compression.
Extra pads should be stocked for repairs during longer production runs. Another
alternative is to butt short sections together in a die retainer to replace damaged pads.
Short sections will perform like one piece if properly retained.
Heat buildup due to internal friction (hysteresis effect) is a common cause of prema-
ture failure of urethane. The amount of heat generated is a directed is a direct function
of effective strokes per hour and/or degree of deflection per stroke. Thus, in select-
ing urethane springe or pressure pads, it is important to minimise the percentage of
deflection for longer life, particulary when exceeding 700 strokes per minute.
For high-speed applications (12,000 strokes per hour), select the size urethane that
will provide the least deflection, or no greater than 15 per cent of total height. For
intermitted operation (700 effective strokes per hour), it is safe to select a size which
will have as much as 25 per cent deflection.
For short runs or slow-speed operations (200 effective stokes per hour when long life
is not important), the maximum deflection of 25 per cent can be exceeded. There is
no bottom possition as in steel coil springs.
Urethanes can generally withstand temperatures up to 120°C, although they may
soften before that point and lose some load – bearing capacity. However, upon
cooling, they will return to their original physical characteristics. Urethanes will
withstand temperature down to –55°C. If stored in cold, they should be brought to
room temperature before being used.
The lubricated or dry condition of a load-bearing area is another factor that affects the
stress-strain relationship. For urethanne compressed between parallel plates, there
is a tendency for this lateral movement, a lubricated surface offers essentially none.
If extremely high pressures are requied, laterally. While a clean, dry loaded surface
752 Tool Design

offers essentially none. If extremely high pressures are required, lateral movement
can be prevented by bonding the urethane to metal with double-faced tape with a
commercial urethane adhesive.

Urethane-die designs and application The die design shown on the following
pages illustrate the simplicity of urethane tooling. However, urethane is not limited to
these specific uses; the drawing merely represent successful design principles. The
versatillity of this material is limitted only by the tool engineer’s imagination.

Urethane forming die Typical press-brake applications with a urethane die pad
are shown in Fig. 11.18. V forming is shown at a. The U-forming die at b utilises an
adjustable die retainer containing removable spacers along the sides to confine die
pads of various sizes. Instead of a machined air channel (as shown in a), deflectors
bars are used to create the necessary air space for proper stress relief and deflection
control of the flat urethane die pad when compressed by the punch. End caps should
be used to confine the urethane. This basic design can accommodate a wide range
of V bonds and other forms.
Figure 11.18d shows a radius-forming die retainer fabricated with a steel insert design
permits the greater bottom pressure needed for forming ribs without excessive strain
on the pad. It is an ideal setup for long runs.
A similar adaption is shown at c. Here, deflector bar in place of the machined insert
are positioned under the pad. The bars are graduated in size to control the deflection
of the urethane and thus help to form the difficult bend.
Figure 11.18e shows how difficult forms can be made proveding sharp bends at
concave sections are not required (definition is better with lighter gauges). Again,
deflector bars beneath the die pad provide air space to permit greater penetration
into the pad.
An unusual adaption of the urethane pad is shown in Fig. 11.18f, where the pad is cast
or machined to shape because greater side pressures are required. Compensation
for springback must be provided for in this type of design. A similar applications is
shown at g. In some applications, a preshaped pad produces such great side pres-
sures that side embossing can be accomplished.
Difficult bending applications can be solved by using steel inserts, as shown in Fig.
11.18h. The inserts are formed to the radius of the punch minus the thickness of
the urethane pad. A thin urethane wear pad is also used to protect the die pad. This
accncmical design eliminates the necessity for a more costly, thicker die pad; also less
tonnage is required.
Figure 11.18i shows urethane wipe-down dies. The set-up shown can easily be
installed in a die set for punch-press applications.
Figure 11.18j and k shows the use of a steel bar on the pressure pad to maintain
flatness of the workpiece. The same set-up can be used for a wide range of metal
gauges by shimming behind the forming blocks as shown at j. The wedge release is
optional. A softer grade of urethane can be used for the pressure pad in this type of
set-up.
Using Plastics as Tooling Materials 753

Fig. 11.18 Urethene forming dies: (a) V-forming (b) U-forming (c) Radius-forming (d) radius
forming (with steel insert) (e) Compound bends (f) Overforming (g) Special
side-pressure forming (h) Special shapes (i) Wipe-down die (j) U-forming
(k) Flange-forming (Di-Acro Division, Houdaille Industries, Inc.)

Urethane pressure pads Oil-resistant urethane has significant advantages over


springs in pressure-pad tooling. Much greater pressure per squre inch can be achieved
754 Tool Design

with less deflection, and therefore less space is required in the die. With the proper
selection of grade, the service life of urethane can equal or exceed that of springs.
The choice of grade depends upon the pressure required. Manufactures’ specifica-
tions give this information (see Table 11.2) as a general guide; however, tests may be
necessary to determine the most suitable grade.
Application of urethane as pressure pads is shown in Fig. 11.19. At a, a standard
urethane pressure pad is machined to fit around a punch. This is an easy, economi-
cal method of constructing a durable, oil-resistant, high-pressure stripping pad. The
application in Fig. 11.19b is a quick, economical way of making a stripping pad for a
short-run blanking die. Urethane strips are fastened (with adhesive or double-faced
tape) around the punch and in the die cavity to function as reliable strippers. Strippers

Table 11.2 Selection of urethane tooling materials


K-Prene* Hardness Tensile Maximum Advantages Disadvantages
grade durometer strength deleftion, %
MPa

K-420 80A 17.92 35 Less to nnage required; Low Tear Strangth;


greater deflection low cut resistance;
permits deeper pene- low load-bearing
tration (best for (pressure) capacity;
lighter gauges) fair forming definition
K-100 90A 31.02 30 Good tear strength;
good cut resistance;
good load-bearing
capicity best abradion-
wear resistance
(recommended for
mosr applications)
K-167 95A 35.85 25 Besr tear strength; Requires more tonnage
best cut resistance; to deflect
gives best definition
with less penetration;
high load-bearing
capacity (best for
heavier gauges)
K-315† 79D* 75.83 5 Good abrasion and Limited to applications
wear resistance; such as forming,
excellent machinability clamping, fixture
offers greatest rigidity blocks, and draw dies
yet will deflect; greatest
load-bearing capacity
*Trade name
†Grade L-315 is much harder, and is meadured on durometer D scale
SOURCE: Di-Acro Division, Houdaille Industries, Inc.
Using Plastics as Tooling Materials 755

of this type are generally used with template, steel-rule, and convection blanking
dies.

Fig. 11.19 Urethane stripping pad for (a) piercing die and (b) blanking die (Di-Acro Divi-
sion, Houdaille Industries, Inc.)

Urethane embossing dies In embossing, urethane performs like a male die when
compressed, forcing the metal into cavties with uniform pressure throughout. Since
only the female die is made, great savings in machining time and costs are possible.
Sometimes it is advantageous to make the male in steel and to use urethane as the
female die.
When designing urethane embossing dies, the surface area of the workpiece, its
thickness, and extent of forming will dictate proper die-pad size and also whether a
die retainer is required. In general, pad thickness should be a minium of three times
the depth retainer is required. In most applications, die pads 6 to 18 mm thick are
used. Thinner pads usually do not require a retainer (because of less pressure loss);
however, greater tonnage will be needed. Thicker pads will create increased side
pressure, thus requiring the use of a strong die retainer.
The use of urethane in embossing operations is shown in Fig. 11.20. The rib-em-
bossing die at a is made by machining rib cavities into the steel punch. This is the
only major tooling work required for this design. A retainer is not needed for this die
because the urethane provides enough pressure to force the metal into the shallow
cavity. The die pad is bonded or tapped to the sub plate.
Fig 11.20b shows the use of a retainer on an embossing die. Note the method by
which the urethane pad is mounted in the retainer. The pad may also be held by
adhesive or double-faced tape. A retainer is usually necessary to reproduce deeper
patterns and sharp detail.
Fig.11.20c shows a special cast or machine-to-shape urethane pad used for a rather
difficult embossing operation. This design provides the necessary side pressure with-
out overdeflection.

Urethane bulging dies In bulging applications, urethane offers much longer service
life over rubber, plus exceptional economy over costly expanding or shrinking steel
punches, Its high tensile strength is achieved without any filler, and it is not affected
by ozone or oil. Typical applications are shown in Fig. 11.21.
756 Tool Design

Fig. 11.20 Urethane embossing dies: (a) Rib-embossing die (b) Rib-and-cavity embossing
die (c) Embossing a drawn shell (Di-Acro Division, Houdaille Industries, Inc.)

Fig. 11.21 Urethane bulging dies: (a) Bulging punch for predrawn pan (b) Bulging nipple
into tubing (c) Bulging shell in split die (Di-Acro Division, Houdaille Industries,
Inc.
Using Plastics as Tooling Materials 757

Figure 11.21a shows the forming of a predrawn rectangular frying-pan cover. The
displacement of the urethane punch forces the metal outward into the desired shape.
The amount of bulge is determine by the depth of the stroke. When the ram returns,
the elastomer resumes its normal shape and the bulge part can easily be removed.
A nipple is bulged in round tubing in Fig. 11.21b. The round bulging punch is deflected
by a plunger and then bulges into the open cavity. Even concolute (coiled or spiraled)
shapes can be bulged in this manner without marring the surface finish.
The urethane punch in Fig. 11.21c is held by a stripper bolt and bulges in proportion
to the depth of stroke. The top alignment plastes must be split in order to remove the
workpiece.

Urethane draw dies Urethane draw dies have two important advantages:
(i) Prefinished metals can be drawn without marking, and (ii) Tooling costs are
significantly reduced for low-and medium-production runs. A urethane draw die may
also accomodate a range of metal thicknesses (maximum 20 gauge) but should be
used only for parts having an outside metal thicknesses (maximum 20 gauge) but
should be used only for parts having an outside flange. Die clearances are less critical
than with steel dies, yet dimensional accutacy in the formed part is maintained (see
Fig. 11.22a). To create desirable side pressures, the urethane pad is machined to an
ID slightly smaller than the OD of the drawn shell.
Urethane pressure pads are used to eliminate springs or air cushions. The design
shown in Fig. 11.22a could be reversed, with the die pad mounted on the bed. In this
case, the drawn part could drop through the die. When drawing tapered walls, com-
pression of the urethane also minimises time the metal is being drawn in the air, thus
helping to avoid wrinkling.

Fig. 11.22 Urethane draw dies: (a) Urethane draw die with steel punch (b) Steel draw die
with urethane punch (c) All-steal draw die with urethane wear pad (Di-Acro
Division, Houdaille Industries, Inc.)
758 Tool Design

Figure 11.22b shows a steel draw die utilising a urethane punch. Here the urethane
functions as both punch and pressure ring (in a retainer). Under compression, the
blank is drawn up into the cavity. However, sheets of urethane can be used as the
punch instead of a solid pad. In addition to greater forming efficiency, these individual
sheet can be interchanged as they wear. For example, the worn top sheet can be
placed on the centre of the urethane punch and thus greatly extend the service life. In
addition, a steal insert can be beneath the punch to concentrate pressures properly
and relieve strain.
A urethane wear pad can also greatly improve performance in convention all- steel
draw dies (see Fig. 11.22c).It compensates for variations in stock thickness, elimi-
nates draw marks, absorbs wear, and provides additional side pressures for more
accurate forming. The steel die is constructed with extra clearance to permit the wear
pad to be drawn along with extra clearance per side between the punch and die
should accommodate the thickness of the metal blank plus approximately 80 per cent
of the thickness of the wear pad prevents direct contact with the metal being drawn.

Urethane inside – flanging dies This type of tooling makes use of a conventional
die with an undersized steel punch plus a urethane wear pad (see Fig. 11.23). The
wear pad, having approximately the same opening as the piece part, is “extruded”
along with the part. Tooling of this type has good abrasion and wear resistance, plus
resilliency, which protects the surface finish. The wear pad also provides additional
side pressure, which compensates for any stock-thickness variations. Clearance
per side between the punch and die should equal the thickness of the metal being
extruded plus approximately 80 per cent of the thickness of the wear pad. The ure-
thane pressure pad provides uniform hold down pressure.

Fig. 11.23 Urethene inside flanging die (Di-Acro Division, Houdaille Industries, Inc.)

Roll forming with urethane Precision roll forming


with virtually no flat spots and no surface marring can
be done in a conventional punch-type rolling machine
(see Fig. 11.24). A slab or sheet of urethane is sim-
ply passed along under the metal sheet to be rolled.
The amount of pressure exerted the top roll causes a
corresponding deflection of the urethane, thus accu-
rately forming the metal. Various radii can be formed Fig. 11.24 Roll forming with
with the same urethane material by adjusting the urethane (Di-Acro
downward roll pressure roll pressure. The third roll Division, Houdaille
is unnecessary and out-of-place. See Table 11.2 for Industries, Inc.)
selection of urethane for this type of application.
Using Plastics as Tooling Materials 759

Urethane wear pads The use of urethane wear pads has been suggested for most
of the foregoing die designs. In each instance, the wear pad serves as a protective
buffer between the metal blank and die pad to extend pad life. It also helps prevent
surface marring (see Fig. 11.25a).
Figure 11.25b and c shows additional applications of urethane wear pads. In b, a
urethane wear pad is attached to one land of all- steel die.It flows along with the punch
and workpiece to protect against surface marring. In addition, the pad also permits
a tighter fit between the punch and die when bottoming. Bottoming pressures are
equalised for improved bending definition. Clearance per side should be approximately
80 per cent of the thickness of the wear pad used.
Figure 11.25c shows a coneventional all- steel die design utillising a urethane wear
pad bonded to the bottom of the spring-loaded stripper plate. This prevents marring
the workpieces.

Fig. 11.25 The use of urethane wear pads: (a) Wear pad for urethane dies (b) Wear pad for
steel dies (c) Wear pad for stripping plate (Di- Acro Division, Houdaille Industries,
Inc.)

Urethane clamping jaws Urethane can for clamping jaws by either machining
or casting the contours into a solid block (see Fig. 11.26). This method of tooling
is especially useful when holding workpiece that vary slightly in size because the
urethane will deflect to accommodate the variations. Complete surface protection is
also possible when machining superfinished workpiece.

Fig. 11.26 Urethane clamping jaws: (a) Urethane chuck chuke jaws (b) Urethane fixture
jaws (Di-Acro Division, Houdaille Industries, Inc.)
760 Tool Design

Urethane fixtures Urethane makes excellent stops and locating pads in fixtures
that hold prefinished workpiece that cannot be scratched or marred. Figure 11.27
shows an assembly fixture that positions a washer or dryer conrrol panel.

Fig. 11.27 Urethane assembly fixture (Di-Acro Division, Housdaille Industrie Inc.)

11.6 CALCULATING FORCES FOR URETHANE


PRESSURE PADS
The following data and information were developed by the Di-Acro Division of
Houdaille Industries during a two year period. Testing was controlled conditions and
the data reflect the variation of E vs. the shape factor for three urethanes with dry and
lubricated surfaces. The curves in Fig. 11.28 represnt a statistical average of the test
results and are offered as a guide to help the enginner predict his forces, size and
grade of urethane required, or per cent deflection.
When selecting urethane springs, strippers, and pressure pads, major consideration
should be given to two factor: the modulus of elasticity and the shape factor.
The shape factor accounts for urethane blocks or cylinders building at their sides
when under a compressive load. Increasing the area that is free to bulge permits
greater vertical displacement or the same displacement or for the same displacement
requires greater force.
The concept of shape factor is numerically defined as the area of one loaded surface
divided by the total area of the unloaded surfaces that are free to bulge. Dimensionally,
this may be written
lw
SF = ________ for rectangular blocks
2t (l + w)
d
___ for solid disks and cylinders
4h
where l = length
w = width
t = thickness
d = diameter
h = height
These equations are limited to piece which have parallel loading faces and pieces
whose thickness is not more than twice the smallest lateral dimension.
Using Plastics as Tooling Materials 761

The modulus of elasticity E is defined as the force per unit area (stress) divided by the
percentage of the change in height (strain), or
F/A
E = _____
DH/Ht
where F = force
A = area
DH = change in height
Ht = total height
For many of the common engineering materials, such as steels, E is a specific value
that remains constant within the elastic range of the material. With urethane, however,
the E value changes as the shape factor changes. Further, its value also changes with
each specific compound (see the curve in Fig. 11.28).

Fig. 11.28 Variation of the moudulus with shape factor. Inset curve for greater than 20
per cent deflection. (Di-Acro Division, Houdaille Industries, Inc.)

In solving problems, the engineer should first determine the shape factor using the
formula previously introduced and assume a particular size of urethane pressure pad.
Having once determined the shape factor, the engineer can then select the appropri-
ate grade of urethane and come horizontally across on the chart to find E. With the
modulus of elasticity known, basic formula of E to solve for either force or percentage
of deflection. That is,
762 Tool Design

DH
F = ___ AE* = (% deflection) AE*
Ht
DH F
% deflection = ___ = ____
Ht AE*

Solved Examples

A forming job on a 150 KN press requires a pressure pad which


Example 11.1 will exert no less than 60 KN at bottom of stroke. It cannot exert
more than 110 KN of force since 40 KN will be used for the
actual forming. Die design has fixed the height of the pressure
pad as 50 mm, the amount of deflection as 10 mm, and the space
available for the pad as 100 by 180 mm. For the blank holding
near bend lines, force must be maintained near the periphery
of the pressure plate (same as die opening). Thus, the pressure
pad can be no less than 50 by 100 mm. Determine the size and
grade of the urethane pressure pad.
Solution: Assuming that the largest pad that will fit in to the given die opening, allowing
space for bulging, is 75 by 150 mm, then the shape factor will be
lw 75 × 150
SF = _______ = ______________ = 0.5
2t (l + w) 2 × 50 (75 + 150)
The modulus of elasticity, E will be
F/A
E = _____
DH/Ht
where DH = change in height, Ht = total height
Given, maximum force exerted, Fmax = 110 KN
Therefore,
110000/(75 × 150)
E = _______________ = 48.89 N/mm2
10/50
The deflection in this case = 10/50 = 0.2, which is 20%.
Since the deflection is 20% no correction of E is required. On the graph (Fig. 11.28)
we find that the junction point E and SF falls around the curve for K-167. Therefore,
we should select K-167 and check that the force does not fall below the 60 KN mini-
mum. Referring to Table 11.2, we get for K-167, E = 35.85 MPa = 35.85 N/mm2.
DH 10
F = ___ EA = ___ × 35.85 × 75 × 150 = 80.662 KN.
Ht 50
The force acting at the bottom of the stroke is 80 KN which is above 60 KN and well
within 110 KN limit.

* For deflections greater than 20 per cent, E must be modified by the multiplier derived from the small curve in
the upper left-hand corner. For deflections equal to or smaller than 20 per cent, this multiplier is 1.
Using Plastics as Tooling Materials 763

If a change in die design permitted a pad 40 mm thick, what


Example 11.2 grade of urethane should be used?

Solution: The shape factor will be


lw 75 × 150
SF = ________ = ______________ = 0.625
2t (l + w) 2 × 40 (75 + 150)
The modulus of elasticity, E will be
F/A 110000/(75 × 150)
E = _____ = _______________ = 39.11 N/mm2
DH/Ht 10/40
where DH = change in height, Ht = total height
10
The deflection = ___ = 0.25, which is 25%.
40
Since the deflection is more than 20%, we find the multiplier for E from the curve in
the upper left hand corner. For 25% this is about 1.09. Therefore,
E = 1.09 × 39.11 = 42.629 N/mm2.
On the graph we find that the new junction point for E and SF falls between K-100
dry and K-167 dry. Thus we choose K-100. Referring to Table 11.2, we get for K-100,
E = 31.02 MPa = 31.02 N/mm2.
Again checking the bottom force,
DH 10
F = ___ EA = ___ × 31.02 × 75 × 150 = 87.243 KN
Ht 40
The force acting at the bottom of the stroke is 87.243 KN which is above 60 KN and
well within 110 KN limit.

Summary

Modern manufacturing forced industry to look for new engineering materials. Plastics
are being used as a engineering materials, even as tooling materials. This chapter
discussed the different types of plastic material, their applications, and construction
of plastic tooling
• Plastics can be formed to desired shape by use of master. They can be worked
at room temperature and are made to take shape of the workpiece by various
techniques of laminating, pasting and casting.
• Various types of plastics that can be used as tooling materials are phenolics,
polyesters, epoxies and urethane compounds etc.
• Phenolics are used for such tools as stretch-form dies and large master fix-
tures, etc.
• Polyster is used for making large moulds.
• Epoxy plastics tools are widely used in stretch dies, draw and form dies,
hydraulic press dies, drop-hammer dies, checking fixtures, assembly fixtures,
drill jigs and machining fixtures, etc.
• Urethane is mainly used in metal forming dies.
764 Tool Design

Questions
1. What is the major advantage of plastics as a tooling material?
2. What are the disadvantages of phenolics as tooling materials?
3. What is the disadvantage of polyesters as tooling materials?
4. What are the primary plastic tooling materials in use today?
5. What is a thermosetting resin?
6. What accounts for the high strength of epoxy tooling?
7. Why are urethane referred to as elastoplastics?
8. How are epoxy tool altered or repaired?
9. Generally speaking, what determines the choice between epoxy and meytal
tooling?
10. Why type of tools can be produced by the use of epoxy resins?
11. What types of tools can be produced by the use of epoxy resins?
12. What are the four basic techniques used in the construction of plastic tools?
13. How is a laminated-plastic tool constructed?
14. When are laminated-plastic tool constructed?
15. What construction gives the tool the best dimensional stability?
16. What is the approximate thickness of each layer of glass cloth and laminating
resin?
17. What is a glass-paste mixture, and how is it used?
18. Why should the lay-up procedure be limited to 18 layer at any one time?
19. Why is the framework fastened to a laminated tool immediately after laminating
and before removing the tool from the model?
20. Why is it best to make the framework of a laminated tool of a material exactly
like the tool facing?
21. How is a surface-cast tool constructed?
22. What types of tool are generally made by the surface-cast method?
23. What is the thickness of the surface-cast facing?
24. What thrsee methods are used in surface casting?
25. What materials are generally used to construct the core for surface-cast
tools?
26. What is one of the greatest disadvantage of mass-cast tools?
27. What is the advantage of adding fillers for large mass castings?
28. What is meant by the epoxy-allow method producing high-production plastic
dies?
29. What is one of the major advantages of paste plastics?
30. What construction technique is used when no master model or pattern is
available?
31. What construction by trowelling and splining?
32. What types of drill bushing are used when constructing drill jigs with epoxy
resins?
33. What are the advantages of urethane plastics over rubber as a tooling
material?
Using Plastics as Tooling Materials 765

34. What is the major difference between metal forming tooling and urethane form-
ing tooling?
35. How is it possible to direct the forces of a urethane die pad?
36. Why should sharp punch ends be avoided in the design of urethane forming
tooling?
37. What precautions should be taken in selecting urethane tooling for high-speed
applications?
38. Why are the advantages of urethane draw dies?
39. What are the advantages of urethane of urethane draw dies?
40. Why are urethane clamping jaws especially useful when holding workpieces
that vary in size?

Design Problems
1. Design a milling fixture to hold the bearing cap shown in Fig. 11.29a. The bot-
tom surface is to be face-milled as indicated by the finish symbols. Assume that
the workpiece is a rough casting without a previous machining operation. Nest
the workpiece in epoxy resin. Use conventional clamping methods.
2. Design a milling fixture to hold the workpiece shown in Fig. 11.29b. The bottom
surface is to face-milling as indicated by the finish symbols. Assume that the
workpiece is a rough casting without a previous machining operation. Nest the
workpiece in epoxy resin. Use conventional clamping methods.
3. Design a milling fixture to hold the workpiece shown in Fig. 11.29c. The slots
are to be milled on a vertical milling machine using end-milling cutters. Assume
that the workpiece is a rough casting without a pervious machining operation.
Nest the workpiece in epoxy resin. Use conventional clamping methods.
4. Design a drill jig to drill the holes in the workpiece shown in Fig. 11.29d. Assume
the workpiece in epoxy resin. Use conventional clamping methods.
5. Design a milling fixture to hole the workpiece shown in Fig.11.29e. The bottom
surface is to be face-milled as indicated by the finish symbols. Assume that the
workpiece is a rough casting without a previous machining operation. Nest the
workpiece in epoxy resin. Using conventional clamping methods.
6. Design a bulging die for the workpiece shown in Fig.11.30a. Select urethane
grade according to Fig. 11.28.
7. Design a bulging die for the workpiece shown in Fig. 11.30b. The workpiece
is to be formed from a drawn cup that has been formed in a previous drawing
operation. Select urethane grade according to Fig. 11.28.
8. The workpiece in Fig. 11.31a. is an aluminium casting that has been cored to
reduce weight. The core used in the casting process is a shell core produced
on a shell-core machine. The moulds on the shell- core machine are to be pro-
duced on a shell-core machine. The moulds on the shell- core machine are to
be produced on a duplicating mill similar to those described in Chap. 2. Design
a duplicating master to be used on the duplicating mill for the purpose of sink-
ing the moulds for the shell-core machine. Use epoxy resin as the material to
duplicating master.
9. Design a form die for the workpiece shown in Fig. 11.31b. The number of work-
pieces to be produced is 150. Use plastic tooling methods to reduce tooling
766 Tool Design

cost. Use existing press in school shop or manufacturer’s brochure for press
dimensions.

Fig. 11.29 Plastic tooling problems

10. Design a urethane forming die to form the workpiece shown in Fig. 10.41a.
11. Design a urethane bending die to bend the workpiece shown in Fig.10.39b.
12. Design a urethane forming die to form the workpiece shown in Fig. 10.39e.
Using Plastics as Tooling Materials 767

Fig. 11.30 Urethane tooling problems

Fig. 11.31 Plastic tooling problems


768 Tool Design

References
Books
• “Metal Forming with K-prene.” Di-Acro Division, Houdaille Industries, Inc Lake City, Minm.
1970.
• Wilson, F W (ed.): “Tooling for Aircraft and Missile Manufacture,” McGraw-Hill, New York,
1964.
• **: “ Plastic Tooling for Aircraft and Manufacturing Handbook,” Prentice-Hall, Englewood, Cliffs,
N J, 1965.
Catalogues
• Devcon Corporation, Danvers, Mass.
• Di-Acro Division, Houdaille Industries, Inc., Lake City, Minm.
• Furnace Plastics Plastics, Inc., Lansing, Mich.
• Pennasalt Chemicals Corp., Philadelphia.
• Ren Plastics, Inc., Lansing, Mich.
• United States Gypsum, Industrial Sales Division, Chicago.
TOOL DESIGN FOR NUMER-
ICALLY CONTROLLED 12
MACHINE TOOLS

12.1 INTRODUCTION
The greatest metal working advance during this century is the development of
numerically controlled machine tools, heralded as “the beginning of a second industrial
revolution.” Since its introduction, more articles have been published in technical
journals on numerical control than any other subject related to the metalworking field.
Numerical control, sometimes referred to as symbolic control, is here to stay. It is a
major step toward job-shop automation, and progressive machining industries are
making good use of it.

12.2 THE NEED FOR NUMERICAL CONTROL


All machine tools require some type of control system for their operations and feed
movements. The most versatile control system has been man, but he is far from being
the most reliable. He has a limited memory and may be readily influenced by his
environment. His speed is slow and erratic, and his ability to repeat dimensions and
co-ordinate more than one dimension is extremely limited.
Control systems have been developed to overcome the many shortcomings of man as
a control system. Instructions to the machine concerning feed movements, positioning,
cycling, and sequencing have been built into the machine in the form of cams, stops,
timers, and other mechanical and electrical devices that are manually established by
the operation before each particular job. This type of control is restricted to parts of
similar configuration. Automatic screw machines, automatic lathes, manufacturing-
type milling machines, and automatic grinding machines are examples of machine
tools employing this type of control; they are often referred to as fixed-program
machine tools. The use of a battery of fixed-program machine tools connected by
transfer mechanisms has resulted in a system capable of producing a large number of
parts. However, the expense of developing and constructing such a system prohibits
its use for small-lot production. For example, it is excellent for the automotive industry,
where thousands of identical parts are produced, but it is not suitable for the aircraft
industry, where in comparison to the auto industry, relatively few parts are produced.
In short-run and individualised production, it becomes necessary to develop variable-
program machine tools.
The need for variable-program machine tools has led to the development of numerical
control. The main contributor to its development has been the aircraft industry because
frequent design changes and small-lot production made it impossible to use automated
production methods as used in industries where long production schedules and the
manufacture of repetitive parts are common.
The production of aircraft for the United States government required the use of
conventional manual- and tracer-controlled machine tools. That type of toolroom and
job-shop production was expensive and could not be relied upon to provide adequate
output of aircraft components. As a result, early developments in numerical control
770 Tool Design

were financed by the United States government. The first numerically controlled milling
machine came about through an Air Force development contract with Massachusetts
Institute of Technology. This machine is generally considered as the first successful
numerical-control application for machine tools, and thus established the concept of
numerical control for industrial use.
Research and development have advanced significantly since the introduction
of numerical control. Reliability has been increased, and the cost of numerically
controlled machine tools has been reduced. Today, even the smallest machine shops
are able to afford the cost of commercial applications.

12.3 BASIC EXPLANATION OF NUMERICAL


CONTROL
A numerically controlled (N/C) machine tool is one that is controlled by numbers.
Instructions for the N/C machine are written in numerical form. These numerical
instructions are written (programmed) in advance and stored on a suitable medium,
which is usually 25 mm wide punched tape, although magnetic tape and punched cards
have been used. The numeric instructions are fed into a main control unit, where they
are stored, interpreted, and changed into signals that are understood by the machine
tool. The main control unit consists of a system of electronic interpreting devices, and
in general is the director of all machine operations. The coded instructions can control
the machine slide positions, spindle speeds, direction of spindle rotation, amount
of feed, direction of feed movements, flow of coolant, and sequence of machining
operations and can select the cutting tool for each operation. On some sophisticated
machining operations and can select the cutting tool for each operation. On some
sophisticated machines, the coded instructions may be co-ordinated to control the
machine axis movements to machine complicated three-dimensional surfaces. The
machine operator’s role is reduced to threading the tape, starting the cycle, loading
and unloading the workpiece, and observing the machine operation. The instructions
to the machine can easily be changed by replacing the roll of tape on the main control
unit with another, similar to the way a movie-projector operator changes a roll of
movie film.

12.4 NUMERICAL CONTROL SYSTEMS IN


USE TODAY
Two basic numerical control systems are in use in modern industry today, a positioning
system (point-to-point) and a contouring system (continuous-path). Positioning
systems are used to locate a point, or a series of points, by moving independently in
two dimensions called x and y dimensions. The path of their motions is not important
as long as they meet at the desired point. Upon reaching the desired point, a third
vertical motion called the z motion can be brought into use to provide depth control.
For example, a N/C drilling machine will position the work at a single specific point.
The drill will advance to drill a hole to the proper depth, withdraw when completed,
and then reposition to the next hole and start the cycle once again. This operation
will continue until the hole pattern is completed. Figure 12.1 shows a numerically
controlled drilling machine. Other machine tools using this type of control system are
jig borers, turret lathes, boring mills, engine lathes, and machining centres.
Tool Design for Numerically Controlled Machine Tools 771

In contouring or continuous-path systems, the path the tool will follow through the
work is important. The x and y dimensions, or the x, y, and z dimensions, are needed
to determine the desired curve or form and must be closely co-ordinated. The
dimensions must be synchronised by the system after being fed into it as individual
dimensions. The curve or form is generated by a series of minute overlapping straight
lines or parabolas. None of these lines or parabolas is longer than the total dimensional
tolerance on both sides of the line or cutter path. By releasing the proportional signal
values of the x, y and z dimensions in each of a series of tiny time intervals, each
interval produces a short straight line or parabola. The sum in continuous systems, as
the thousands of points needed for contouring prohibit manual programming. Figure
12.2 shows the use of a continuous-path milling machine.

Fig. 12.1 A numerically controlled Fig. 12.2 Using continuous-path milling


drilling machine (Colt In- machine (Cincinnati Mila-
dustries, Pratt & Whitney cron, Inc.)
Machine Tool Division)

The machining-centre concept Early efforts in the development of numerically


controlled machine tools were concentrated on machines that were to do a job
previously impossible or extremely difficult or too expensive to do by conventional
means. As the N/C concept developed and materialised, it was recognised that N/C
could be very easily adapted to conventional machines. Machine-tool manufacturers
realised that simple, low-cost, and high-volume machines controlled by N/C had a large
commercial market potential. This type of machine was developed and marketed.
Conventional machine tools equipped with N/C relieve the operator of many planning
and positioning responsibilities; however, many of the original limitations still exist.
It continues to be necessary to reposition the workpiece manually to machine all
surfaces, manually change the tooling, and in many cases, manually select the feeds
and speeds. Excess time for material handling and set-up continues to be a problem.
The limitations of N/C conventional machine tools led to the development of the N/C
machining centre. This machine has the ability to finish parts in a single setting. The
workpiece is located on an index table to provide access to more than one surface.
Three linear axes are available, and feed rates and spindle speeds are selected by
772 Tool Design

tape. The machine has a magazine


for storing the necessary tools and a
tool-changing arm that transfers the
tool from the magazine to the machine
spindle. Some machining centres
contain a workpiece shuttle or transfer
device that allows the operator to
load one workpiece while machining
operations are being performed on
another. Figures 12.3 and 12.4 show
two different machining centres. A
tool changer is shown in operation
in Fig. 12.5. The machining centre in
Fig. 12.6 contains a workpiece shuttle Fig. 12.3 A numerically controlled machining
mechanism. centre (Kearny and Trecker Corpo-
ration)

Fig. 12.4 A numerically controlled machining centre (The Sundstrand Corporation)

Fig. 12.5 The tool changer of a numerically controlled machine tool (The Kearney and
Trecker Corporation)
Tool Design for Numerically Controlled Machine Tools 773

Fig. 12.6 A numerically controlled machining centre utilising a workpiece shuttle


mechanism (The Kearney and Trecker Corporation)

The advantages of a N/C machining centre are quite obvious. After the initial set-up,
additional set-up time is eliminated. Only one work-holding fixture is necessary
because the transfer to other machines has been eliminated. Set-up time is reduced
for the same reason. Less time is needed for the change of feeds and speeds, tool
selection and insertion, and indexing of the workpiece. Machining time can also
be predicted because the human element has largely been eliminated during the
machining process. A more accurate part is produced because of the elimination of
transfer error.

The advantages and disadvantages of numerical control Numerical control has


been a major step toward automating the job-shop and short-run production. It has
done for the small shop what Detroit automation has done for mass production. The
advantages are many, and a few have already been discussed. However, there are
cases where conventional machine tools are better suited. There is still a definite
place for conventional machine tools. For example, a manufacturing firm that
produces occasional large castings or weldments in small lots may be better off using
a standard radial drill. The initial cost of a N/C machine large enough to handle this
type of production cannot be justified unless it is kept busy with the type of production
for which it was designed. On the other hand, a lower-cost conventional radial drill
could handle the occasional large castings on weldments and in the meantime be
used for the production of smaller parts. The lower initial cost of the conventional drill
may be justified when producing parts of a large variety of piece sizes.
The criteria for justifying the purchase of a numerically controlled machine tool are
many. There are those who say that economic justification on paper is impossible,
yet these same people are buying N/C equipment and vastly increasing their profits.
Many companies get into numerical control by buying one of the smaller and less
expensive machines as an experimental venture. The results are sometimes fantastic.
There have been cases where the machine has payed for itself within three or four
months. The advantages and disadvantages of N/C are summarised in Table 12.1.
774 Tool Design

Table 12.1 Impact of numerically controlled machine tools


Field Advantage Disadvantages
Engineering Reduced data; closer tolerances; quick Lofted or numerical definitions
design-change capability; all parts alike; require NC drafting machine
higher contouring complexity to check designs
Manufacturing Low tool cost; faster tool and first-part Limited maximum production
delivery; flexible quantities; less rate; no alternate method if
specialised operator skills; higher designed and tooled for NC;
operator efficiency; less material mastered tooling out; no
handling “work around” capability
Production control Reliable production estimates and yields; Less machine-planning
fixed flow time flexibility
Methods and Time-study and incentive systems not Methods improvements, part of
standardisation required to attain efficiency tape program, not close to
floor conditions
Inspection and More uniform quality; “on cycle” More division of responsibility
quality control inspection; random defects of NC of product quality
operation are usually gross and easy
to see
Plant engineering Fewer machines to house and service NC equipment more complex;
special training and skills are
required to maintain NC; spare
parts inventory higher;
inoperative = 100% down
Purchasing When possible, subcontract Material supply must be
reproducibility and control is maintained; limited sources of
excellent subcontracting
Accounting Better const controls; better cost Overhead and indirect costs
information; less scrap higher; higher capital
investment
Sales Firmer real cost and producibility Less chance to improve a “too
control; better capability to sell low” quote
SOURCE: The Bendix Corporation, Kansas City Division.

The material thus far in this chapter has dealt almost entirely with numerical
control. How does numerical control affect the tool designer? Are the fixtures used
on numerically controlled machines any different from those used on conventional
machines? What about cutting tools? Should not drills used with conventional machine
tools be suitable for use on a N/C machining centre?
The answers to the above questions are both yes and no. The basic fundamentals
of fixture design do apply to the design of fixtures for N/C machines. Likewise, the
cutting tools used on N/C machines are similar or may be the same as those used
on conventional machine tools. However, the introduction of numerical control has
placed new demands on conventional fixturing methods and cutting tools. In the same
manner, it has placed new demands on the machine tool itself. For example, the
human operator may be able to cope with worn or nonrigid conventional machine
tools. A skilled machine operator may lean on one end of the machine table during
the cut to prevent one or two thousandths of unwanted taper. Not so with numerically
Tool Design for Numerically Controlled Machine Tools 775

controlled machine tools. The machine itself must produce the actual cut. Machine
tools must be built especially for N/C, and as a result, a much more rugged, rigid, and
friction-free machine has evolved.
Twist drills are a good example of cutting tools. Conventional drilling machines employ
a drill jig containing hardened bushings to guide the drill. The drill jig and the bushings
are eliminated on N/C drilling machines. The work is positioned accurately under the
spindle, and the drill must start itself. Conventional drill geometry may allow the drill
point to wander and cause inaccurate hole location. It is necessary to use a drill-point
geometry that will not allow drill wander.
It is the purpose of this chapter to investigate the needs of tooling as applied to
numerically controlled machine tools. It is intended to supplement the information
provided in previous chapters.

12.5 FIXTURE DESIGN FOR NUMERICALLY


CONTROLLED MACHINE TOOLS
Numerically controlled machine tools are classified as a universal machine tool; in
that they are capable of producing a wide variety of workpieces. This is due, for the
most part, to the fact that they are a variable-programmed machine. With this thought
in mind, consider fixtures for N/C machines. The use of a mill fixture designed for one
shape and size of workpiece, as generally used on a conventional production milling
machine, would contradict one of the underlying principles of N/C. N/C should reduce
the amount of necessary tooling. It is quite apparent that this goal cannot be reached if
a fixture is designed and constructed for every different workpiece produced. Fixtures
of this type are expensive to manufacture and maintain, they take up storage space
when not in use, and are obsolete if the workpiece is hanged or eliminated.
The answer is to design tooling that is universal. Fixtures should be designed to hold
more than one size of workpiece. The fixture should be able to do one job and then be
rearranged to fit a variety of other workpieces. This is not to say that all workpieces can
be held in a universal fixture, as complicated shapes may require a special fixture. The
goal, however, should be to hold workpieces in universal fixtures whenever possible.

Strap clamps and heel blocks Common work-holding devices on a standard machine
tool are strap clamps and heel blocks. They may be located in almost any position on
the table, and the only restriction on location is the position of the T slots on the table
they are universal in nature, in that they can hold a wide variety of workpieces. With
the addition of suitable stops mounted on the machine table to locate the workpiece,
they make an excellent universal fixture for a N/C machine tool. They are especially
well suited for N/C drilling machines. Figures 12.7 and 12.8 show typical set-ups. This
type of fixturing is suitable for small-lot production, the number of parts ranging from
5 to 50.
The experienced tool designer may be reluctant to accept this type of fixturing. He
must be reminded that part of the concept of N/C economy is simple tooling and a
minimum of tooling. Tool designers and toolmakers may have to be reeducated to the
fact that N/C requirements are different.
776 Tool Design

Locators

Parallels

Fig. 12.7 Using strap clamps and heel blocks to hold flange plate on numerically controlled
drilling machine

Perhaps a better approach to fixturing for


N/C with strap clamps and heel blocks would
be to design and build the heel block and
strap clamp into one unit, as shown in Fig.
12.9. In this case, the typical heel block is
replaced with a screw which enables an
operator to adjust for varying workpiece
heights. A spring holds the strap clamp in
position while the operator is replacing the
workpiece. The gooseneck design of the
strap keeps the total assembly low to give
additional tool clearance.
Work supports and locators may also be
incorporated into one unit, as shown in Fig.
12.10. They should be heat-treated for wear Fig. 12.8 Using strap clamps and heel
resistance and brought to size by machine blocks to hold workpieces on
grinding. The T nuts and cap screws in the a numerically controlled ma-
assembly can be purchased as standard chining centre (The Kearney
items. and Trecker Corporation)
The use of the strap clamp, support, and locator assemblies previously described is
essentially the same as conventional fixture design, except that the machine table is
used as the fixture baseplate. The main difference between the clamps, supports, and
location is that the assemblies can be relocated to accommodate different sizes of
workpiece. The fundamentals of good fixture design still apply.
This type of tooling is furnished as standard tooling for some N/C machines. Figure
12.11 shows standard tooling available for the Pratt and Whitney Tape-O-Matic drilling
machine. The workpiece supports, locators, and clamps are incorporated into separate
nesting blocks that can be mounted at varying distances from each other. Two of the
clamps are quick-acting for increase speed, while the other clamps are held by hex
nuts to provide positive clamping. Note that the individual workpiece-support blocks
Tool Design for Numerically Controlled Machine Tools 777

are held into position by magnets and therefore can be located in any position on the
machine table. Figure 12.12 shows the workpiece in place on the fixture component.

Fig. 12.9 Strap-clamp assembly for holding work on numerically controlled machines

Fig. 12.10 Workpiece locators and supports for use on numerically controlled machines

A device for locating the spindle of a N/C drilling machine directly over the zero point
of square or rectangular workpiece is shown in Fig. 12.13. In this case, the corner of
the workpiece was designated as the origin, or zero point, by the programmer. The
locating device, referred to by the manufacturer as an Accuset, is used in conjunction
with the nesting blocks shown in Fig. 12.11. The Accuset locator is placed against
the two corner nesting blocks. A setting plug is placed in the hole of the Accuset,
and both dial indicators are set to read zero. The setting plug is removed and a
proving bar of the same diameter as the setting plug is placed in the machine spindle.
778 Tool Design

Fig. 12.11 Using clamp nesting-block sets on Fig. 12.12 The workpiece in position on
a numerically controlled drilling standard fixturing compo-
machine (Colt Industries, Pratt & nents (Courtesy of Kansas
Whitney Machine Tool Division) State College of Pittsburg)

Fig. 12.13 Using the Accuset zero locator Fig. 12.14 Standard milling-machine vise
(Colt Industries, pratt & Whitney used on a numerically controlled
Machine Tool Division) drilling machine (Courtesy of
Kansas State College of Pitts-
burg)

The table is jogged to a position approximately over the hold on the Accuset. The
spindle is lowered until the proving bar contacts the indicators. The table is then
jogged in low speed until both dial indicators again read zero. The x = 0 and y = 0 axis
positions are then locked in the machine control and the zero position is established.
When the square or rectangle workpiece is placed in the nesting blocks, the corner
will be on zero position.

Standard vises and chucks Many workpieces machined on N/C mach-ines can be
conveniently held in standard milling-machine vises. The main requirement is that the
workpiece be located in the same position each time. An adjustable locator may be
Tool Design for Numerically Controlled Machine Tools 779

attached to a standard milling-machine vise, as shown in Fig. 12.14. The locator is


designed to allow adjustment in two directions in order to accommodate various sizes
of wrokpiece. This type of fixture is limited to smaller workpieces; however, two vises
may be located side by side on the machine table. This arrangement allows workpiece
loading and unloading while the machine is completing another part.
Adjustable vises are available as standard items, which allow for a larger range of
workpiece sizes (see Fig. 12.15). Note that the vise is composed of three components
that are butted in various positions on the machine table. One component provides
jaw movement while the other two serve as stationary locators. The three components
are keyed to slots in the machine table.

Fig. 12.15 An adjustable vise designed especially for numerically controlled machine
applications (Sundstrand Machine Tool)

Three-and four-jaw chucks mounted


on a flat plate are all that are required
to hold many parts to be machined on
N/C machine tools (see Fig. 12.16).
Again two or more chuck units may
be located on the machine table in
order to alternate workpiece loading,
unloading, and machining.

Grid pattern on base plates An


auxiliary baseplate containing a grid
pattern of holes is especially suited
for N/C drilling machines (see Fig.
12.17). The holes are usually on Fig. 12.16 Three-jaw chuck mounted on the
25 mm centre with every other hole table of vertical-spindle numerically
tapped to take hold-down studs or controlled machine (Sundstrand
screws. The remaining holes are jig- Machine Tool)
bored to 5 microns accuracy and are
used to receive dowel pins for workpiece locators. The grid plate is bolted directly to
the machine table. The grid pattern of dowel-pin and stud holes provides numerous
locating combinations for positioning pins and studs and may be used for an infinite
variety of workpieces.
Referring again to Fig. 12.17, we see that a set-up hole in noted for a zero reference
point. This places the grid pattern in the upper right-hand quadrant, and all dimensions
can be expressed as positive terms. In use, the N/C machine table is manually
positioned until the spindle is directly over the set-up hole. The relationship of spindle
780 Tool Design

to set-up hole is established with a dial indicator mounted in the spindle. The x = 0
and y = 0 axis positions are then locked, and the machine is ready for operation. All
table movements will be made in reference to the set-up point. The workpiece is then
located on the auxiliary grid plate in reference to the set-up point.

Fig. 12.17 Grid plate used for location of workpieces on numerically


controlled drilling machine

The use of a universal grid plate enables the programmer to specify the location
of the workpiece. He is able to determine the location by sketching the workpiece
outline over a full-scale master layout of the gird plate. Reference holes or surfaces
on the workpiece ar established in relation to one or more master dowel holes in the
grid plate. The method and location of clamps are also determined at this time. The
sequence of operations, pattern of machining, table motion, etc., are all in relation to
the original set-up point.
Odd-shaped parts that cannot be held readily on the grid plate can be held in special
auxiliary fixtures built on small bases containing holes that match those on the
grid pate. The auxiliary fixture is then pinned and clamped to the grid plate, and all
references are made to the original setup point (see Fig. 12.18).
Special grid plates are sometimes made directly on the N/C machine. In this case,
the toolrooom is by-passed completely. A cold-rolled steel plate is mounted on the
machine table, usually raised to allow drill clearance and to allow chips to pass through
holes in the grid plate. The locations of the dowel and tapped holes are programmed
on tape and are drilled and bored directly on the N/C machine.
When the workpiece is small enough to allow two pieces to be mounted on the grid
plate at one time, it is possible to machine one piece while the other is being removed,
depending upon the type of cutting tool or number of operations and tool changes. The
programmer specifies the location of both workpieces with reference to a common
setup point on the grid plate.
One disadvantage to the grid system is that chips tend to accumulate in holes. This
can be partially overcome by keeping all unused holes plugged, especially threaded
Tool Design for Numerically Controlled Machine Tools 781

ones. Loos-fitting plugs that are easily removed can be made to prevent chips from
accumulating in dowel holes.

Fig. 12.18 Auxiliary fixture to be mounted on master grid plate

Several manufacturers of N/C machine


tools offer the grid system and building-
block principle as standard tooling.
Figure 12.19 shows the grid plate
and components offered by Brown
& Sharpe Mfg. Co. It consists of a
baseplate which has a standard hole
pattern with known centre distances
between standard holes and sets of
knees of various sizes which also have
the same grid pattern as the base plate.
The knees can be placed accurately on
Fig. 12.19 The Brown and Sharpe grid system
any part of the baseplate because the
of fixturing (The Brown & Sharpe
holes are a very close matched pattern.
Mfg. Co.)
Adjustable and angular stops which fit
the grid are designed to make it possible to position any part on the grid baseplate in
any necessary position. Adjustable jacks screw into the top of the baseplate. A tooling
system of this type will accept almost any shape of workpiece that has to be held.
The tool designer or part programmer, depending upon the policy of the company,
lays out the workpiece set-up on a grid drawing, giving locations of stops, clamps,
782 Tool Design

workpiece, etc. Locations are easily specified because holes running across the grid
plate are marked A, B, C, etc., and the holes running down are marked 1, 2, 3, etc. A
record is kept, and regardless of when the job comes to the machine-six months or
six years later-the set-up remains the same. The programmer in turn programs the
machining operation in accordance with the workpiece location. A tool-set-up sheet
is sent to the machine operator along with the programmed and punched tape. The
operator then assembles the tooling components on the grid pate as called for on the
tool-set-up sheet and places them in the holes as marked, such as plate stop X in
hole A1.
The Brown and Sharpe grid system simplifies tool design because the basic
components are a known increment. All the tool designer has to do is determine
where the part will be placed on the baseplate, and then he can set stops or jacks as
they are needed to complete the set-up.
For example, assume that a ring has to be set-up and located from the inside diameter.
This is simple to accomplish using the Brown and Sharp grid system. The part can be
set in an accurate position in the absolute centre of the machine rotary table using two
45° stops and one plain stop. If an operator had to accomplish this without a special
fixture, it would take at least 3 hr of measuring and clamping before the machine
was ready for operation. On the other hand, if a fixture was made to hold the part in
position, the fixture would cost many man-hours and dollars and still would just be
another single-purpose tool. It would require upkeep and storage, and by the time the
fixture was complete, a tool designer could draw set-ups of many parts on the grid
system.

Fixturing for N/C machining centres The Sundstrand Machine Tool Company has
standardised fixturing composed of modular components based on the grid principle.
However, in this case, the workpiece locators are located approximately on the grid
pattern and then precision-machined under tape control. Dowel pins are not used. The
procedure is a decided advantage where extreme locational accuracy is require.
The Sundstrand system is offered for use on the Sundstrand five-axis Omnimil.
The Omnimil is a complete numerically controlled machining centre that will mill,
bore, ream, tap and contour-mill any exposed face of the workpiece at any angle.
It incorporates an automatic tool changer and pallet changer which permits bench
loading of parts to reduce machine idle time (see Fig. 12.20).

Fig. 12.20 The Sundstrand model OM-3 Omnimil with automatic tool changer and pallet
change (Sundstrand Machine Tool )
Tool Design for Numerically Controlled Machine Tools 783

Figure 12.21 shows a standard modular


fixture that is mounted on the Sundstrand
N/C machine table. Set-up of the part is a
simple procedure. Where required, part
locators are bolted to the modular fixture
and precision-machined, as shown in
Fig. 12.22.
Sundstrand standard fixture plates permit
bench loading of parts to reduce machine
idle time (see Fig. 12.23). Lots of four of
the parts shown were processed in 7.4 Fig. 12.21 Standard modular fixture for
hr, including set-up, as opposed to 36 hr Sundstrand Omnimil (Sund-
by a previous method. strand Machine Tool)

The Kearney ad Trecker Corporation has developed standard universal fixtures and
free-tooling concepts principally for use on the Kearney and Trecker Milwaukee-Matic
machining centres. The following discussion will be concerned with universal fixturing
as applied to the Milwaukee-Matics; however, the concept may be applied to other
horizontal-spindle machines.

Fig. 12.22 Use of modular fixture Fig. 12.23 Use of standard fixture plates to
components (Sundstrand permit bench loading (Sundstrand
Machine Tool) Machine Tool)
784 Tool Design

A model II Milwaukee-Matic is shown in Fig. 12.24. This machine, like all true machining
centres, has the ability to finish parts at a single setting. It incorporates a tool-change
mechanism which inserts a new tool in the machine spindle and simultaneously
returns the used tool to the magazine when called for on tape. It has an index table,
providing access to more than one side of the workpiece. An automatic shuttle-
loading mechanism allows automatic reciprocal operation between two different parts.
Loading, set-up, and change-over are performed with machine downtime.

Fig. 12.24 A Kearney and trecker Model II Milwaukee-Matic machining centre (Kearney
and Trecker Corporation)

Generally speaking, most workpieces machined on N/C controlled Milwaukee-Matics


and similar horizontal-spindle machines fall into one of three groups:
1. Workpieces that are positioned on a flat (or relatively flat) side and are machined
about the periphery.
2. Workpieces that are standing on edge and are machined form either one face
of opposite faces.
3. Workpieces that can be supported in a conventional three-jaw lathe chuck
mounted on a vertical angle plate.
Examining group 1, where the part is oriented on its back and machined about the
periphery, some interesting facts may be noted:
1. Most fixtures incorporate some sort of riser blocks which position the part
some distance above the machine pallet in order to eliminate possible interfer-
ence between the spindle nose and the pallet, particularly at some 45° index
increment.
2. The fixture, exclusive of the riser blocks, is basically much like a lathe faceplate
lying flat and providing facilities for clamping studs.
Ordinarily, in this type of fixture design, the tool designer will include these riser blocks
and clamping studs to suit the particular part under consideration. As a result, we find
a condition as shown in Fig. 12.25. These examples show three different fixtures with
several common features:
1. All three include a baseplate attached to the machine pallet.
2. All three include riser blocks (two different heights are shown.)
3. All three include a top plate with provisions for accepting clamp studs.
Tool Design for Numerically Controlled Machine Tools 785

Fig. 12.25 Examples of workpieces oriented on their back and machined about the periphery
(The Kearney and Trecker Corporation)

Since all three include similar components used for identical purposes, it is logical to
assume that one universal pedestal base could be used in place of the three fixtures.
The pedestal base would incorporate provisions for locating it on the pallet. Since this
base would have a predetermined height, for example, 125 mm, a standard y-axis
reference zero point would exist for programming purposes.
The top plate of this pedestal would have a pattern of tapped holes in it which would
enable it to be used much like a lathe faceplate with straps, studs, fixed locators,
etc. It could also accept the base of a simple plate-type fixture for each part (see Fig.
12.26).
The advantages in adopting this universal
fixture are many. Some examples are
the elimination of risers from individual
fixtures, the elimination of accurate jig
boring or milling of individual fixtures for
pallet-locating keys, and the reduction of
fixture storage space.
It is sometimes required to index the
work pallet through various degrees of
position to accommodate the machining
of some parts; therefore, it may be
necessary to make the top member plate
of the pedestal circular. This plate, as
well as all the other members, should be Fig. 12.26 Universal pedestal base fixture
heavy stock not less than 25 mm thick (The Kearney and Trecker Cor-
to eliminate deflection while machining. poration)
786 Tool Design

Universal pedestals should be provided with top plates of various sizes. As a


suggestion they could be 230, 305, and 460 mm in diameter.
Equally interesting facts can be obtained from a study of fixtures which orient the part
on an edge. Fixtures of this type are generally constructed as follows:
1. A horizontal baseplate is used with three jig-bored locating holes to accom-
modate pins for locating on the pallet.
2. Attached firmly to the base is a vertical plate which carries a pattern of straps.
clamps, locators, etc., to position and secure the workpiece.
3. Gusset plates are arranged and secured so as to support the vertical plates.
Bacause of this duplication of similar components, it is logical that one unit incorporating
a universal base and vertical plate and gusset could replace the corresponding
components on all the individual fixtures. As a result, the fixtures themselves would
consist of a plain vertical plate and associated clamping devices. This plate member
would be supported by the universal fixture and attached to it. The principles outlined
in Fig. 12.27 may be used as a guide or pattern for design. Figure 12.28 shows the
Kearney and Trecker standard universal fixture of this type.

Fig. 12.27 Guide and pattern for design of universal vertical fixture
(Kearney and Trecker Corporation)

If access to the rear of the part is required,


access holes can be drilled or bored through
the vertical plate under tape using the proper
co-ordinates from the part-machining cycle.
If the work to be performed is confined to
one face of the part, it is possible to load two
fixtures for two different parts back to back
on opposite sides of the frame uprights for
more advantageous cycle performance. The
part configuration may be such as to require
repositioning several times during the cycle.
The back-to-back arrangement is convenient
in this case. The part can be machined and
loaded progressively from side to side of the
fixture with no interruption in the machining
cycle of the other part. The Kearney and Fig. 12.28 Universal standard vertical
Trecker approach to fixturing offers two fixture (Kearney and Trecker
advantages: Corporation)
Tool Design for Numerically Controlled Machine Tools 787

1. Fixture storage: By eliminating the base members as well as riser blocks, the
fixtures occupy less cubic storage space. Because of the change in physical
shape they can be stored on edge rather than on the locating surface. As a
results the locating surfaces are less likely to be damaged.
2. Fixture Handling: Thanks to the reduction in weight, they are easier to han-
dle. Fixtures can be moved into and out of storage racks by hand in most
cases, thus reducing the necessity for elaborate materials-handling equipment.
Because the basic vertical frame remains attached to the pallet, there is a dis-
tinct reduction in the possibility of damage to the pallet surface and precision-
locating T slots when setting up and removing the fixture.
It will be noted that a rather large group of parts which are located on edge and
could be machined with the subplate large group of parts which are located on edge
and could be machined with the subplate discussed above may also be machined
under a different and still more inexpensive set of conditions. These conditions are
characterised by the commonly called Tinker Toy or Erector Set fixturing concept.
These terms refer to the basic principle of utilising a standard set of components or
subassemblies and arranging them into a wide variety or combinations to meet the
fixture requirements. The cost of the individual standard components may appear
rather high. This is due to the cost of engineering design to make them universal.
However, considered as being shared among a group of parts over a period of time,
the cost per part is reduced tremendously.
For example, a single-purpose fixture may cost $500. This same fixture assembled
form a universal set could cost as much as $2000 but would be arranged to handle 10
or more different parts. Consequently, the basic cost of the universal fixture is divided
by the number of different arrangements it will be used for. In the example, 10 is the
number used. This would reduce the fixture cost to only $200 for each part instead of
$500. An example of this approach in fixturing is shown in Fig. 12.29.

Fig. 12.29 The Kearney and Trecker Tinker Toy or Erector Set fixturing concept (Kearney
and Trecker Corporation)
788 Tool Design

The Milwaukee-Matic and other


machining centres may be equipped
with a shuttle system. Downtime for
set-up exists only when the pallets are
transferred to and from the machining
centre. At all other times, the machine
is producing parts. It is only necessary
to provide a program for the part in
advance of the work requirement. Figure
12.30 shows a view of the Milwaukee-
Matic shuttle system.
Conventionally, planning machine
movements are always done by the
machine operator. This practise is not Fig. 12.30 The shuttle mechanism of a
always readily isolated or identified, but Milwaukee-Matic machining
it does exist. Heretofore, it has been an centre (The Kearney and Trecker
practise to ignore nonrecurring small- Corporation)
lot (five pieces or less) and one-piece-
one-time runs as a Milwaukee-Matic work load potential. This is due to disregarding
non-identified or recognised set-up time existing with conventional manufacturing
methods. As a result, too much emphasis has been placed on the relatively larger lot
sizes, which do not occur very frequently, and therefore the work load for the machine
will appear low.
A contributing factor has been the popular misconception that an individually designed
fixture is required for each part produced on a tape-controlled machine. The total cost
of the fixture is applied to the job again and again, and the immediate conclusion is
that part production cost is excessive unless a large lot is run.
Best suited for the one-time or one-shot jobs would be a typed of “free” tooling. For
descriptive purposes consider these free tooling definitions: (i) all free tooling can be
called universal tooling, but (ii) universal tooling is not necessarily free tooling. Also for
descriptive purposes, free-tooled set-up is defined as “an assortment, not necessarily
hardened and ground, of locators, supports, and clamping members so arranged on
a pallet as to adequately locate, support, and secure a workpiece for machining. “Any
or all of the locators and/or supports may conceivably be machined in position to suit
this one piece and then be discarded.
The essentials of a free-tooled set-up are few in number but extremely important.
The set-up must be such as to transpose the part-print reference dimensions readily
and accurately to the known machine reference dimensions. The set-up must also be
such as to transpose the cutter dimensions readily and accurately to the part-print
requirements. The feedback in both cases is the inspection which certifies or denies
the part print. It must be remembered that this certification is valid only on the particular
part inspected and is subjected to all errors of inspection and interpretation.
The necessity for making set-ups rapidly has already been discussed. Normally a
free-tooled set-up can be made quite rapidly provided some means of transposing the
part reference to the machine and verification of cutter dimensions is readily available.
One of the most versatile pieces of equipment for this purpose is the Kearney and
Trecker Polycheck, shown in Fig. 12.31. By placing the machine pallet directly on the
Tool Design for Numerically Controlled Machine Tools 789

established and clearances checked for the workpiece, machine, and tools. This may
be done without introducing any setup time into the machine cycle.
An important and desirable fringe benefit can be obtained by using the Polycheck.
Since the actual workpiece is presented to the spindle in exactly the same way it will be
machined, the cutting tools themselves can be inserted into the spindle and the actual
physical tool clearances can be checked. Figure 12.32 shows tool clearance being
checked on a typical part. Tool-setting distances can be planned to take advantage
of minimum tool overhang, which results in improvement in cutting efficiency. This in
turn will result in lower overall cycle time. The same tools can be used when the job is
actually machined on the Milwaukee-Matic.

Fig. 12.31 The Kearney and Trecker Poly- Fig. 12.32 Use of the Kearney and Treck-
check (The Kearney and Trecker er Polycheck (The Kearny and
Corporation) Trecker Corporation)

The initial reference-location coordinates of the part can be transferred directly to the
programming work sheet and used in conjunction with the various part dimensions
called for on the engineering drawing to complete the par program itself. Finally, the
tape can be punched and program type-out verified, and the part is ready to run. Note
that no actual contact with the Milwaukee-Matic is involved. Program refinements in
cutting speeds, feeds, and the number of passes can be made only during actual part
production on the machine.
In some cases, the quantity of parts to be run at any one time may be so low that the
additional time to refine the cutting speeds may not be justified in terms of the saving
in cutting time.
Another type of free-tooled set-up that can be made very rapidly, again using the
Polycheck, is a precision angle plate mounted on the pallet. Black, soft-steel locator
blocks and rest pads are mounted on the angle plate to suit the part configuration.
By using the Polycheck a program can b established to machine these locators and
pads on the Milwaukee-Matic. The tool list and tape, as well as a sketch showing
790 Tool Design

the placement of the pads and locators, are


retained for future use. After using the set-up
to machine the parts, the soft locators are
simply discarded. In the event of rerun, the
set-up can be duplicated by again positioning
the soft blocks and machining then, using the
same tape.
Another type of free-tooled set-up can be
made by utilising a standard milling-machine
vise mounted on an angle plate. Again using
the Polycheck, the most advantageous
location of the vise can be established. A
ground dowel pin can be assembled to a
toolholder and programmed to a convenient
position near the vise jaws and used as
a positive location point for the part. This
location, verified on the Polycheck, is then
used to program the part. If convenient, a Fig. 12.33 Use of standard mill vise
locator can be provided as an x-axis locator. mounted on angle plate (The
Figure 12.33 illustrates an example of this Kearney and Trecker Cor-
kind of set-up. The obvious limitation of this poration)
type of fixture is the size of the vise jaws.
To provide a much greater holding capacity than is available in the standard milling-
machine vise, a universal vertical vise, as shown in Figs 12.34 and 12.35, was
developed. The jaw-opening capacity of this device is readily apparent. The lower,
semifixed locating jaws can be positioned in any one of three horizontal key slots.
The basic semifixed upper-jaw assembly can be positioned so that the movable jaw
can be adjusted to permit securing and releasing the part with a minimum of jaw

Fig. 12.34 Use of universal vertical Fig. 12.35 Use of universal vertical vise
vise (The Kearney and (The Kearney and Trecker
Trecker Corporation) Corporation)
Tool Design for Numerically Controlled Machine Tools 791

movement. All adjustable side bracket, usable on either side on the vise base, can be
positioned to provide a locating end stop for the workpiece. The two-way positioning
of this bracket facilitates locating the stop screw to avoid interference with cutters, the
clearance for this screw can best be verified with the Polycheck.
Figure 12.34 shows two adjustable under the lower jaws to provide additional support
when removing stock from the large joint face. Clearance for these jacks should be
verified on the Polycheck. The vise-jaw inserts used in both of the jobs shown are of
the serrated type. For use with semifinished parts, smooth, hardened, or soft jaws
may be installed.
Figures 12.34 and 12.35 show a left- and a right-hand base. When required, the units
may be placed adjacent to each other to provide greater width capacity. When used
this way, the combined width of the two bases is within the pallet edges and the entire
assembly can be moved with respect to the x axis to provide the best tool “reach”
setting. Tie bars may be used across the top of the assembly to connect the two units
into one rigid assembly.
The previously mentioned three- or four-jaw chucks are relatively inexpensive, and
one may be mounted permanently to a small angle plate. Once the chuck has been
mounted, the basic dimension for x, y, and z programming reference can be determined
and recorded. Where extreme accuracy is required, the chuck may be fitted with soft
jaws and bored in position to suit the job much the same as for second-operation lathe
work or internal grinding. Tool clearances are checked and verified with Polycheck.

12.6 CUTTING TOOLS FOR NUMERICAL


CONTROL
A numerically controlled machine tool is only as good as the cutting tool. Even though
the spindle may be more accurate, the machine is more rigid, and the positioning
system more precise than the equivalent conventional machine tool, the N/C machine
is no more accurate than the cutting tool in the spindle. Cutting tools that may be
suitable for conventional machines may not be applicable for N/C machines because
N/C machines exhibit greater dependability of the N/C machine.
The use of N/C machines has simplified or eliminated elaborate jigs or fixtures.
This may simplify production, but it will increase the need for exact and rigid cutting
tools. Cutting-tool inaccuracy or lack of rigidity may be overcome with conventional
machines because the jig or fixture will, in part, guide and support the cutting tool and
produce greater accuracy.
The following discussion applies to specific types of cutting tools used on N/C
machine tools and is more concerned with the selection of available standard tools
than with the design of specific tools, although the modification of standard tools will
be discussed.

Twist drills Generally speaking, production drilling on conventional drilling machines


involves the use of drill jigs containing guide bushings. The drill is positively located
and guided into the correct position of the workpiece. The drill is supported to a degree
after entry, especially where long drill bushings are used, and therefore the tendency
to chatter is reduced.
792 Tool Design

Thus, the drill depends upon the guide busing for location. The drill-point geometry
may be poor, yet satisfactory results are still obtained when suing guide bashings. Drill
concentricity can be slightly off since the guide bushing will have it corrected and is
true enough for satisfactory results.
The use of N/C drilling machines generally eliminates drill guide bushings, and the
drill is no longer dependent. Hole location, size, and quality are determined by the drill
itself, assuming that the positioning system of the N/C machine is functioning properly.
For example, the N/C machine may position the table within 12 mm; however, this is no
assurance that the location of the finished hole will be to this accuracy. The drill may
be running out of true or it may walk slightly as it enters the workpiece. Improper point
to be off-centre. This will result in holes being oversize. The cutting-lip angles may
be unequal and cause one cutting edge to work harder than the other. This causes
torsion strain, bell-mouthed holes, and poor tool life.
Standard drills that are geometrically correct may cause problems when used
on N/C drilling machines. Point geometry may have to be modified to produce
satisfactory results. Standard drills may have to be replaced with special drills for N/C
applications.
Consider the standard twist drill. It is probably the most common metal-removal tool in
modern industry. It is capable of removing more volume of metal per minute than any
other rotating cutting tool of equivalent diameter. However, the basic design of the
standard drill has not changed during the past century. Basically, it contains a shank,
flutes, margin, web, two cutting edges, and a chisel edge. The chisel edge does not
form a true point. Should one end of the chisel edge be slightly higher than the other,
the drill will tend to walk to one side or the other of the desired hole location. This can
be avoided to an extent by first centre-punching the hole location, but this is a hardly
acceptable practise when using N/C drilling machines.
One of the first requirements for drills used on N/C drilling machines is a drill point
that will centre itself in the correct location without walking. The spiral or helical point,
developed a few years ago, has eliminated drill walking to a large degree (see Fig.
12.36). As indicated by the term spiral point, the entire flank surface associated with
each cutting edge is formed as a three-dimensional spiral surface, extending from the
axis of the drill to its periphery. Thus, the drill has a true point because of the way in
which the spiroidal surfaces are ground.

Fig. 12.36 The spiral-point drill (Cincinnati Milacron, Inc.)


Tool Design for Numerically Controlled Machine Tools 793

In order to make the spiral point of practical use, a machine is required which will grind
this geometry accurately and economically. A machine having these qualities, the
Spiropoint drill sharpener is shown in Fig. 12.37. This machine is so designed that the
drill is held stationary while the generating system gyrates around it. Other grinding
machines are on the market that will produce similar drill-point geometry.
A procedure often used to establish drill location when using standard drills is to
precede the drilling operation with a spotting or centre drill (combination drill and
countersink) for each hole in the pattern. Spotting and centre drills are shown in Fig.
12.38. Spotting drills are ground in a manner to produce a nearly true point, and the
drill portion of the centre drill is so small that the chisel edge is reduced to almost a
true point. Both tools are extremely rigid because of their shorter length and heavier
shank. Standard centre drills may have to be held in an extension because the spindle
travel on countersink) for each hole in the pattern. Spotting and centre drills are shown
in Fig. 12.38. Spotting drills are ground in a manner to produce a nearly true point,
and the drill portion of the centre drill is so small that the chisel edge is reduced to
almost a true point. Both tools are extremely rigid because of their shorter length and
heavier shank. Standard centre drills may have to be held in an extension because
the spindle travel on the smaller N/C drill presses may not be long enough to use the
shorter centre drills without moving the position of the drill head.
The spotting or centre-drilling procedure requires an additional sequence of operations
but may be justified because of greater accuracy in hole location. The included angle
of the centre drill or spotting drill should be narrower than the included angle of the
drill that follows it to direct the drill into the centre of the spotted hole. The depth of the

Fig. 12.38 Spotting drill and centre drills


used for accurately staring
Fig. 12.37 The Spiropoint drill sharpener standard twist drills (The Moore
(Cincinnati Milacron, Inc.) special Tool Company)
794 Tool Design

centre drilling or spotting operation should


be slightly over the drill diameter for the
best centreing action (see Fig. 12.39). This
may also eliminate a debarring operation,
as the minor burr produced when the drill
enters the workpieces will be reduced
in size. The small burr produced will be
below the wokrpiece surface and thus
may not have to be removed.
Use of spotting drill may eliminate a
countersinking operation. The spotting
drill is ground to different point angle to
correspond to the included angles of the
countersink. This procedure works well
Fig. 12.39 Relationship of spotting-
for countersinking holes that are to be
or centre-drill diameter to drill
threaded.
diameter
The preceding discussion notes the
importance of rigidity in shanks of spotting and centre drills. Rigidity is also important
in twist drills used on N/C drilling machines. The construction of a standard twist drill
does not lend itself to maximum rigidity. The main body of the drill is cut away to
form drill flutes and leave a tapered web. Drills constraining long flutes lack torsional
rigidity, especially near the drill point, where the web is narrow. Drills tend to wind and
unwind during the cut, and in extreme cases chatter may be provoked by the winding
and unwinding action. This hammering action quickly shortens tool life and reduces
hole quality as well.
Since long flutes affect torsional rigidity, it is apparent that a shorter flute is
advantageous, providing the depth of hole will allow it. One way to get shorter flutes
using standard drills is to use a collet-type holder that allows the drill to be gripped
on the flute portion as shown in Fig. 12.40. This method of holding not only increases
torsional rigidity but increase lateral rigidity as well. A spiral-point drill held by this
method will produce holes with good locational accuracy.
Another way frequently used to reduce the flute length is to use screw-machine-
length drills. This type of drill is standard and was developed for use in screw
machines, where conditions require short drills. Short flutes and short overall length
for maximum rigidity are characteristics of this type of drill. Similar results can be
obtained by shortening a standard-length drill. However, since the web of standard
drills is tapered, shortening the flutes will require thinning the web when the point is
retargeted. The various methods of web thinning are discussed in Chapter 4.
Complete drill symmetry is necessary for optimum performance on N/C drilling
machines. First of all, the drill must be held concentrically. Collet-type drill holders will
provide better concentricity than standard drill chucks and should be used whenever
possible.
The cutting lips and the web of the drill should be concentric for accurate hole sizing.
When the point or web is not concentric, a relative difference in lip height results,
causing the two cutting edges to form chips unevenly. The result is that the drill is
deflected with the result that the hole is oversize, the hole finish is poor and the drill
life is shortened. Thus, if maximum drill performance is to be obtained, high-quality
Tool Design for Numerically Controlled Machine Tools 795

Fig. 12.40 Collet-type holder designed to grip a twist drill on the flute portion (Erickson
Tool Company)

drills should be purchased for use on N/C drilling


machines. Many drill companies are producing
this type of drill to meet the needs of N/C drilling
machines. New manufacturing techniques have
been developed to produce improved point
geometry. Figure 12.41 shows a drill designed
especially for N/C drilling. Its short length
provides maximum rigidity without give, and its
helical point is self-centreing. The flutes are fast
spiral and polished to prevent loading. The drill
markings are electroetched on the shank rather
than stamped. This feature helps to eliminate
concentricity errors produced by chucking Fig. 12.41 A twist drill engineered
against a protrusion caused by stamping on the especially for numeri-
drill shank. cal-control drilling (The
Jarvis Corporation)
Another drill designed for N/C drilling is shown
in Fig. 12.42. It features a double margin combined with short length, special point
geometry, and close tolerances. The double margin increases the bearing surface
and helps to reduce deflection. More accurate hole sizes are possible, and in some
cases, a reaming operation can be eliminated on some materials.
In summary, the requirements for a twist drill to be used on N/C drilling machines
are extreme rigidity of the drill body, good concentricity throughout, geometry that
will break the chip, and a point geometry that will not walk as the drill enters the
workpiece. A drill with these characteristics will produce on-location holes, frequently
796 Tool Design

with reamed-hole tolerances in certain


materials. This type of drill must be used if the
full accuracy potential of a N/C positioning
system is to be utilised. However, it must be
remembered that locational accuracy is not
always the reason for drilling workpieces
on a N/C drilling machine. The reason may
be the increased drilling speed, and it may
not matter if the hole pattern or hole size is
within 0.8 mm. A typical example would be
cover plates for coal washing equipment.
Here standard jobber’s-length twist drills
would be satisfactory in most cases. The only Fig. 12.42 A double-margin drill de-
advantage in using the more accurate and signed for use on numerically
rigid twist drill would be that the feed could be controlled drilling machines.
substantially increased. Note the short length for
added rigidity. (Mohawk
Spade drills The discussion thus far has tools, Inc.)
been confined to smaller twist-drill sizes.
When it is necessary to drill holes larger than 25 mm diameter on N/C drilling
machines, many users have found spade-drills satisfactory and eonomical. Figure
12.43 shows a typical spade drill.

Fig. 12.43 Spade Drill (Muskegon Tool Industries)

Spade drills will produce straight and accurate Chip-breaker


holes from the solid. The drill point has a grooves
small dead centre, which minimises end thrust
and drill walking when the tool enters the
workpiece. The stubby bar design increases
rigidity and resistance to torque. Chip breakers
are incorporated into the drill geometry to
break the chips as they are formed. Figure
Chipcurl
12.44 shows the basic geometry of the spade
drill.
Fig. 12.44 Basic geometry of spade
The initial cost of spade drills is considerably drill (Muskegon Tool In-
lower than the standard twist drills required to dustries)
cover the same range of sizes. For example,
a kit of eight bars offered by one manufacturer will accommodate 256 different size
cutters. Each cutter may be reground several times.

Milling cutters Generally speaking, the cutting speed, feed rate, and depth of cut
the higher on N/C milling machines or machining centres. For this reason, the trend
has been toward heavier and stiffer milling cutters made of the premium grade of
Tool Design for Numerically Controlled Machine Tools 797

cutting-tool materials. Solid carbide tools have had increased use because of the
inherent rigid stability of tungsten carbide.
One of the major problems with milling cutters for N/C machines is that programs are
written for specific cutter diameters. Therefore, when the cutter becomes dull and is
reground to a smaller size, it is no longer suitable for the program. This is especially
true of end mills. The solution to this problem has been approached in several ways.
Some companies use reground cutters for roughing operations and reserve the new
cutters for finish work only. This practise allows longer cutter life for the finishing cutter
that has the programmed diameter. Some control systems are available with a cutter-
offset option, which allows cutter compensation by the operator. The operator is able
to correct manually an offset error caused by a changed diameter. The operator is able
to correct manually an offset error caused by a changed diameter. Other companies
use new cutters of programmed diameter only, and when they become dull, they are
retargeted and used on conventional milling machines.
Another solution has been to make
additional tapes that allow for the
reduction in the diameter of the
cutter. However, in may cases, it is
less expensive to replace the cutter
than to go to the added expense of
programming a new tape.
One of the common methods of
eliminating the mill-diameter problem
is by using insert-type milling cutters
(see Figs 12.45 and 12.46). In this
case, the cutters are used until they
become dull. The inserted teeth,
usually held in the cutter body by
serrations, are extended by one Fig. 12.45 Inserted-tooth face-milling cutter
serration and then reground to the (The Lovejoy Tool Company)
original cutter diameter. This type of
cutter is available only in sizes larger
than 25 mm.
An end mill developed especially
for N/C machining is shown in Fig.
12.47. The tool is of solid carbide and
consists of reusable 2410 MPa high-
strength special die grade carbide
shank on which a replaceable carbide
cutting-grade ring is brazed. After
concluding a predetermined number
of resharpenings, the carbide cutting-
grade ring section is removed, and a Fig. 12.46 Insert-type milling cutters (Sonnet
new ring is brazed in its place. The Tool and Manufacturing Company)
deflection of this tool is held to an
absolute minimum because of the inherent rigid stability of tungsten carbide.
798 Tool Design

Fig. 12.47 Solid-carbide end mills designed especially for numerically controlled machining
(The Metal Removal Company)

Throwaway insert-type milling cutters are often used on N/C milling machines,
especially for roughing cuts. Some users avoid this type of mill for finish cuts because
they claim that surface finish is sacrificed by slight irregularities in tooth position
and length. Other users claim to get satisfactory surface finishes using throwaway
insert-type milling cutters providing proper techniques and cutter designs are used.
For example, the face-milling cutters shown in Figs 12.48 and 12.49 feature an
adjustable locator cam pin that permits axial adjustment of the insert. Face runout
can be controlled to tenths of a thousandth, providing the time and care are taken to
adjust the insert properly when quality surface finishes are required. The face mill in
Fig. 12.49 also features a fine pitch that allows faster feeds without increased chip
loads. This allows full utilisation of the higher feed rates available on heavy-duty N/C
machining centres.

Fig. 12.48 Coarse-pitch throwaway insert- Fig. 12.49 Fine-pitch throwaway insert-
type face-milling cutter (VR/ type face-milling cutter (VR/
Wesson Company) Wesson Company)

Throwaway insert end-milling cutters are shown in Fig. 12.50. Generally speaking, this
type of end mill is limited to the heavier machines because of shock of the intermittent
straight-flute cut, especially with a negative-rake cutter. This tendency is reduced
with the positive-rake cutter because of the added cutting efficiency provided by the
positive rake angles.

Taps Generally speaking, tapping on a N/C machine is no different from tapping


on conventional machines. In many instances, tapping on N/C machines is easier
because the positioning accuracy is near perfect and the machines spindle, tap
holder, tap, and workpiece are in precise alignment.
Tool Design for Numerically Controlled Machine Tools 799

(a) (b)

Fig. 12.50 Throwaway insert end-milling cutters: (a) Negative rake (b) Positive rake (VR/
Wesson Company)

One problem encountered in N/C tapping is that the operator may not be watching
the tapping operation as closely as he would on a conventional drill press or tapping
machine. It is therefore necessary to select taps that prevent chip accumulation in
the flutes. On through holes a “gun” or spiral-point tap should be used, shown in
Fig. 12.51. The spiral point creates a shearing action during use and projects the chips
ahead of the tap. It may be used on blind holes, providing there is sufficient depth
beyond the threads to accumulate chips. The flutes of this type of tap are shallower
than in regular taps, as flutes are not needed for chip passage. The result is a thicker
cross section of the tap body with greater strength.

Fig. 12.51 A gun or spiral-point tap (United-Greenfield Corporation)

For blind holes in soft and ductile materials, a spiral-fluted tap should be used to help
eliminate flute clogging (see Fig. 12.52). The spiral flutes have an auger effect and
tend to lift the chip out of the threaded hole. Spiral-fluted bottoming taps also help
reduces the chip load per tooth because of the high helix angle.

Fig. 12.52 A spiral-fluted tap (United-Greenfield Corporation)

Many times when tapping under N/C, it is desirable for the tap to extend from the
toolholder, the same distance as the tap drills. This is especially true on N/C machines
that do not have tap dept control in the z axis and rely on limit switches for depth
control. Because of the limited number of depth-control limit switches, it is sometimes
necessary to use the same limit switch for both tapping and drilling operations when
large numbers of different holes are contained in the workpiece. In this case, an
extension or pulley tap may be used as shown in Fig. 12.53.

Fig. 12.53 An extension or pulley tap (United-Greenfield Corporation)


800 Tool Design

The selection of the tapping head or tap driver is quite important when tapping under
N/C. Figure 12.54 shows a compensating tension and compression tap driver that is
used on N/C machines that do not have provisions for adjusting the feed rate to match
the lead of the tap accurately. In use, the tap is allowed to float axially to compensate
for lead error.

Fig. 12.54 Compensating tension and compression tap driver (Scully-Jones Company, sub-
sidiary of The Bendix Corporation)

Many tap drivers have a torque-limiting adjustment as well as axial float. The torque-
limiting adjustment is advantageous when tapping blind holes where the thread must
extend to the bottom of the hole. In use, the feed rate is programmed or manually
selected slightly less than the lead of the tap, depending upon the type of N/C
machine. This causes the tap to feed out of the tap driver (utilising axial float). When
the tap hits the bottom of the hole, the tap stops rotating (utilising torque limit) and
allows the spindle to catch up with it. When the spindle reaches the preset depth, it
will reverse and allow the tap to feed out, rather than breaking the tap when the axial
float is taken up.

12.7 TOOLHOLDING METHODS FOR


NUMERICAL CONTROL
One of the virtues of N/C machine tools is that dead or nonproductive time is reduced
drastically. A machine tool is not making money unless it is making chips. It does little
good to buy a machine tool capable of making heavier cuts at increased feed rates
if it takes the same amount of time for set-up and tool changing. By the same token,
it makes little sense to use the newer cutting-tool materials and increase the cutting
speed to a point where it endangers the operator and still use traditional tool change
and set-up methods. The best way to reduce machining costs today is to reduce
machine downtime.
Early efforts in N/C machine tools largely relieved the operator of planning and
positioning responsibilities. However, he still had to change tools manually for
different operations. When the tool became dull, it was also necessary to change
the tool manually and then adjust it by cut-and-try methods. This is time consuming
and is even worse when the machine must be idle while the adjustments are being
made. Thus, it was inevitable that a new concept in tooling would evolve for use on
N/C machines. This concept has been applied to conventional machines as well.
Basically, it may be divided into two parts: presetting the tools to correct dimensions
and quickly interchangeable tooling. Like the N/C machines themselves, the tooling
may vary from simple and relatively inexpensive to quite elaborate and costly systems.
The purpose of this section is to explore the systems available and to give the tool
designer an idea of how they are used.

Quick-change Tooling Although it is not often thought of as quick-change tooling,


the modern throwaway insert shown in Fig. 12.55 cannot be overlooked. This type of
Tool Design for Numerically Controlled Machine Tools 801

tooling has brought about considerable reduc-


tions in machine downtime for replacement of
dull tools. It is especially useful on N/C lathes
and boring machines. When the cutting edge
becomes dull, it takes only 30 to 45 sec to
index the insert to a new cutting edge. More
importantly, the new cutting edge is in the
same relative position to the workpiece, and
the operator does not have to readjust the tool
to the original reference point. Various types
of throwaway insert tools and their advantages
and use are discussed in detail in Chapter 4. Fig. 12.55 Throwaway insert-type tool-
holder (Kennametal, Inc.)
Spade drills with replaceable blades may
also be classed as quick-change tooling, as the blades can be changed in seconds
(see Fig. 12.56). There may be slight difference in length between new and reground
blades, but this is not important on through holes. When the hole depth is critical, the
machine should be programmed for new blade lengths and the replacement blades
should be new. Reground blades are then used on through holes.

Fig. 12.56 Changing the blade on a spade drill (Muskegon Tool Industries)

The most numerous styles and types of quick-change toolholders are made for single-
spindle N/C machines, such as drilling, jig-boring, and horizontal machines. Each
design offered by the various manufactures has its own unique features, but basically
the majority consists of a master toolholder that is either built into the spindle of the
machine at the time of manufacture or is built into a standard taper that enables it to
be quickly adapted to a standard spindle. Adaptors are then provided to adapt the
various types of cutting tools to the master holder. The adapters usually contain a
taper that fits into the master holder on one end. On the other end, there are provisions
for holding the particular tool, such as collets, tapered sockets, Jacobs tapers for drill
chucks, etc. The master holder incorporates some type of locking device that can be
activated by hand or by a spanner wrench. Some masters may be actuated entirely by
hand, while others always require a wrench. A quarter turn of the wrench is sufficient
to release the tool adapter. Other master holders may be locked and unlocked by
hand for mild machining operation where the cutting forces are constant and always
in the same direction, such as reaming or drilling, but also provide a wrench for more
positive locking where heavy or interrupted cuts are encountered. Figure 12.57 shows
802 Tool Design

the construction of one common quick-change toolholder. The adapters are easily
mounted in the Kwick-Switch holder by simply inserting the tapered shank into the
master holder and turning the locknut a quarter to the right, thus locking the adapter
securely in place. To change tools the locknut is turned to the left, which breaks the
adapter loose from the master holder. A stop pin automatically positions the slots in
an open position so that the adapter can readily be removed. Figure 12.58 shows the
various kinds of adapters available.

Fig. 12.57 Universal Kwick-Switch tool holder (The Universal Engineering Company)

Referring again to Fig. 12.58, note the


rake provided for holding the adapters.
Note also that each tool position is
numbered to identify the tool. A rake
of this type is usually portable so that
it can be moved from the machine
to the toolroom. The adapters and
cutting tools are selected and set at
the toolroom according to instructions
on a tooling work order. They are then
returned to the machine.
In use, the operator selects the
tools according to the process
chart (sometimes referred to as a
manuscript). The tool number on the
process chart corresponds to the Fig. 12.58 Adapters available for Universal
number on the tool rack. When it is Kwick-Switch tooling (Universal
necessary to change tools manually Engineering Company)
during operation, the N/C machine is
directed by tape to move to a tool-change position and stop its sequence in order to
clear the workpiece during manual change. Some machines use a tool-change light
to indicate to the operator that a tool change is necessary. Other machines may be
programmed to show the tool number on a readout system when it is ready for a tool
change. One N/C machine contains individual tool-change lights built into a storage
rack to identify the particular tool.
Tool Design for Numerically Controlled Machine Tools 803

Another type of quick-change toolholder,


manufactured by the DeVlieg Microbore
Division of the DeVlieg Machine Company,
is referred to as the Flash Change toolholder.
The master holder contains an eccentric cam
that engages a groove in the tapered adapter
shank. The cam is rotated with an Allen
wrench, which in turn locates the adapter
radially and locks it in place. A reverse turn of
the wrench unlocks the assembly and ejects
the tools (see Fig. 12.59).
The Flash Change system provides great
versatility in that both the no. 40 and 50 size
taper shank be used in any standard no. 40
or 50 National Machine Tool Builders socket
with either a manual or power drawbar. The
same tooling can then be used on other types
of spindles that have adapted to the Flash
Change system. This versatility becomes very
Fig. 12.59 Flash Change toolholder
important from an economic standpoint when
(DeVlieg Machine Com-
an inventory of tooling is selected to serve a
pany, Microbore Division)
variety of machine spindles.
A quick-change holder manufactured by the Erickson Tool Company is shown in
Fig. 12.60. The master holder contains two lugs that pass through keyways milled
into the adapter flange. This positions the adapter radially. When the locking nut is
turned, its own lugs move over the adapter flange and force it into the taper, locking it
in place (see Fig. 12.61). A unique feature of the Erickson system is that its adapters
are standard milling-machine tapers. Standard driving flanges and drive slots are
used. The result is that Erickson quick-change tool holders will accommodate exiting
adapters with standard milling-machine tapers of corresponding numbers.

Fig. 12.60 Erickson quick-change toolholders (Erickson Tool Company)

The Moore Special Tool Company has used a simple and functional quick-change
tooling system on their jig-boring machines for many years. Basically, it consists of a
taper held in the machine spindle by a fast-lead square thread. In use, a quick spin
of the tool adapter seats the adapter in place, and it is then locked with a wrench.
The reverse procedure removes the tool, and the necessary time for tool change
is approximately 10 sec (see Fig. 12.62). Master adapters are offered to adapt the
Moore taper to other machine tools.
804 Tool Design

Fig. 12.61 Construction of the Erickson quick-change toolholder


(Erickson Tool Company)

When the Moore Special Tool Company introduced their N/C jig-boring machines,
mother quick-change feature was added to increase the efficiency of the N/C jig borer
and similar machines. It is referred to as the Moore wrenchless adapter and is shown
in Fig. 12.63. Wrenchless adapters require about 6 sec for a tool change. They are
not suitable for heavy cutting operations, for which the regular Moore taper shank is
used.

Fig. 12.62 Use of the standard Moore Fig. 12.63 The Moore wrenchless adapter
taper (The Moore Special Tool (The Moore Special Tool Co.)
Company)

The discussion thus far has been concerned with quick-change tooling for N/C
machines with rotating cutting tools. Consider for a moment quick-change toolholders
for N/C lathes. Several Manufacturers offer this type of tooling featuring the same
basic concept. Figure 12.64 shows a popular design of quick-change tool posts. A
master tool post is mounted on the slide of the lathe and locked at the proper angle
by means of a hex flange nut provided at the top of the tool post. A wide range
and type of toolholders (Fig. 12.65) can be easily and quickly mounted on the tool
post by simply turning the independently working lock handle. A knurled nut on each
Tool Design for Numerically Controlled Machine Tools 805

Fig. 12.64 Aloris quick-change tool post Fig. 12.65 Toolholders available for Aloris
(Aloris Tool Company) tool post (Aloris Tool Company)

toolholder provides for exact height adjustments, thus eliminating the use of shims.
Many operators claim that this type of quick-change tooling is faster than a square
turret when several tools are involved. Toolholders can preset to within 25 mm and
repeatability is said to be within 2.5 mm.

12.8 AUTOMATIC TOOL CHANGERS AND TOOL


POSITIONERS
The previous discussion had to do with increasing the speed and efficiency of manual
tool changing. The next logical step in this direction would be to eliminate manual
handling of the tools all together. This has generally been done either by providing a
means of indexing the tools into cutting position, as in the case of a turret drill, or by
providing a tool-change mechanism which removes the tool from a storage magazine
and places it in the machine spindle. Either method adds a considerable amount to
tool cost, but the expense is more than offset by increased machine utilisation.
The advantage of tool changers and tool positioners is that they unerringly place the
tools in position or in the spindle with regular consistency. A good operator may be
able to change tools manually faster than some mechanical tool changers, but he
cannot do so consistently. Management cannot predict the amount of time it will take
for an operator to change tools; however, when using a tool changer or tool positioner,
management can predict tool-change time and therefore retain control of the total
machining process.

Turret-type N/C machines The simplest type of N/C machine with tool positioners
is the turret drilling machine. One of the smaller machines available is shown in Fig.
12.66. This machine is equipped with both power index and automatic cycle, together
with numerical control of table positioning. The automatic cycle includes rapid spindle
transverse to work, power feed control, rapid spindle retraction, automatic index,
adjustable feed rate for tapping, and automatic reversal of the tap at predetermined
depths. It has six turret spindles with a no. 33 Jacobs taper. The time required to index
the turret is 1 sec. It is capable of drilling, tapping, reaming, and counterboring up to
12.5 mm.
806 Tool Design

Fig. 12.66 Turr-E-Tape tape-controlled turret drilling machine


(The Brown & Sharpe Mfg. Co.)

A turret-type N/C machine capable of milling, in addition to drilling operations is


shown in Fig. 12.67. It automatically drills, mills, taps, bores, and counterbores with all
functions selected and controlled by tape. Once the cycle has been started, the entire
sequence of operations is completed without further operator supervision.

Fig. 12.67 The Cintimatic numerically controlled turret drill (Cincinnati Milacron, Inc.)

Turret indexing is by tape command, and any one of eight tools may be selected at
random by tape commands. A maximum of 6 sec is required to change from one tool
to another. Power is delivered only to the tool selected, the others remaining idle.
Tool Design for Numerically Controlled Machine Tools 807

Hole depths are automatically controlled by tape commands, and thus manual depth
stops are eliminated. This feature enables a single tool to drill a variety of depths
during the machining cycle without manual adjustment. The need for presetting tools
is eliminated by use of a tool-length compensator built into the control of the machine.
All that is necessary is to program the hole depth. If a 100 mm depth is desired, for
example, 100 mm is programmed. It is not necessary to program rapid advance and
tool retract.
For example, to compensate for tool length, the operator feeds the turret carrier down
until the tool touches a feeler gauge resting on the workpiece surface. With the tool
resting on a the feeler gauge, he adjusts the no. 1 tool-length compensator controls
on the console for spindle station 1 until the needle on the compensator dial rests on
zero. This adjusts the control circuit for the tool length and work height and establishes
the zero starting point for that tool. The depth is then controlled by tape from this point.
Four tool-length compensator control switches, with adjustments from inches down to
thousandths, are provided for each of the eight spindles.
When the operator completes the tool-length compensating procedure for the no. 1
tool, he retracts the turret and indexes to the next position, using the pendant controls.
The same procedure is repeated until tool-length compensations have been made for
all spindles. The time required is less than 1 min for each spindle.
Where additional tools are needed, the turret machine shown in Fig. 12.68 may be
used. This machine contains 20 vertical quills in a horizontal turret arrangement. Four
of the spindles are heavy-duty milling spindles with automatic quill clamps spaced
90° apart on the turret, although milling, drilling, tapping, and boring may be done on
any spindle. The turret indexes in either direction and permits programming to any
randomly selected tool via the shortest path. The turret indexing speed is 3 sec from
one spindle to an adjacent spindle and 5 sec from one spindle to a spindle spaced
180° away. Figure 12.69 shows a closer view of the turret and spindle mechanism.
The Brugmaster 20T features full three-axis numerical control. Presetting tools and
programming rapid traverse lengths are eliminated by use of random-tool-length

Fig. 12.69 The turret and spindle mecha-


Fig. 12.68 The Burgmaster 20T multiquill nism of the Burgmaster 20T
machining centre (Burgmaster multiquill machining centre
Division of Houdaille Industries, (Burgmaster Division of Hou-
Inc.) daille Industries, Inc.)
808 Tool Design

compensation. This feature also permits the use of random-length tools. Basically,
the random-tool-length compensation feature consists of a zero shift for each spindle,
from the top (or home) position to some intermediate point in the z-axis travel. The
setting is performed by the machine operator during the job set-up, and, depending
upon operator skill, requires less than 1 min per tool.
To perform the random-length tool setting, the operator switches the tape-manual
dial switch to the manual position. The turret is fed down until the top of the tool
contacts a 5 mm feeler block (the feeler can be any thickness, but 5 mm has been
used for this discussion). This will establish the zero height above the workpiece and
the point at which the down feed will change from rapid to feed. The part programmer
has added this 5 mm dimension to his programmed z depth when planning the tape,
and the machine control therefore thinks that the tip of the tool is on the surface of
the workpiece. The operator then switches to dial position from manual and while
depressing the cycle-start push button with one had he advances the z-axis zero
shift switch that corresponds to the spindle with the other hand. The switch contains
a bank of four thumbwheels, each corresponding to digits, tens, hundredths, and
thousandths. Each thumbwheel is advanced individually, commencing with the digit
bank, until the lights on the console switch from increase to decrease. The thumbwheel
is then backed one digit, and the next successive thumbwheel is advanced, thus
zeroing digits, tens, hundredths, and thousandths. When all thumbwheel have been
nulled, the zero for that tool has been accurately shifted from the top of the workpiece
to the point 5 mm above.
A N/C lathe utilising a turret-type tool positioner shown in Fig. 12.70, is known as
the Numeriturn II20. The lathe was specifically designed for numerical control and
features a unique slant-bed design that permits greater bed width with no sacrifice in
space. The sharp slope of the bed moves great volumes of chips quickly and does not
allow hot-chip accumulation under the workpiece or cutting tool.

Fig. 12.70 The Numeriturn II20 numerically controlled lathe (The Lodge and Shipley
Company)

The Numeriturn II20 vertical-indexing turret is shown in Fig. 12.71. It contains six
stations and permits interference-free turning up to the spindle nose for between-
centre work and converts to chucking work without changing the set-up. Turret
indexing is hydraulically actuated by a separate hydraulic system under tape control
and indexing. Repeatability is ±3 seconds of arc. Coolant is individually piped to each
tool for automatic coolant supply when the tool is in use.
Tool Design for Numerically Controlled Machine Tools 809

The turret is provided with six standard


open-slot turning toolholders suitable
for using catalogue-type throwaway
cutting tools. Positive locating surfaces
permit presetting of cutting tools in
the toolholders at the bench. Less
than 2 min is required to change a
toolholder in a turret station. Additional
turning, boring, facing, drilling, and
special toolholders are available for all
machining conditions (see Fig. 12.72).

Automatic tool changers One of the


better known and more popular tool
changers is offered on the Milwaukee-
Matic machining centres. Figure 12.73
shows the tool-changing mechanism
for the Model II Milwaukee-Matic. Tool Fig. 12.71 The Numeriturn II20 vertically
storage is accomplished by a rotary indexing turret showing six preset
magazine for 30 toolholders with tools in position (The Lodge and
tools. With one tool in the spindle, the Shipley Company)
machine has a capacity for 31 different
and automatically transferred tools.

Fig. 12.72 Standard vertical turret tool holders for the Numeriturn II20 numerically con-
trolled lathe (The Lodge and Shipley Company)

The tools are held individually in adjustable holders, which serve to locate the tools
in the spindle and to identify each tool by means of coding rings located at the tool
end of the holder. How tools are secured in the holders varies according to size and
cutting requirements. Figure 12.74 shows a toolholder with coding rings in position.
Eight basic styles of toolholders accommodate te various types of cutting tools. When
more than 31 tools are required to complete a workpiece, stops are programmed
and punched into the tape to allow the operator to remove an replace tools in the
810 Tool Design

rotary magazine. Stops must also be programmed and punched into the tape when
using larger tools having excessive setting length or diameter. This allows manual
tool changing for large tools that would not clear the machine mechanism during
automatic tool change.

Fig. 12.73 Tool-changing mechanism of Model II Milwaukee-Matic (The Kearney and


Trecker Corporation)

Fig. 12.74 Toolholder with coding rings in position (The Kearney


and Trecker Corporation)

The tools may be loaded into the magazine in any sequence. The magazine rotates
by tape command to the proper position to bring the preselected tool into place for
tool change. The tool search (drum rotation) is usually programmed to take place
automatically while the previously selected tool is cutting. When the drum is in
position, the magazine pocket containing the desired tool pivots out, the tool-changer
Tool Design for Numerically Controlled Machine Tools 811

arm swings into tool-change position, and one end of the arm grasps the tool in the
spindle while the opposite end grasps the tool in the magazine. The tool-changer
arm then moves out away from the spindle, removing one tool from the spindle and
the other from the magazine. The arm rotates 180° and brings the selected tool into
position at the spindle nose and the other tool into position at the magazine. The arm
moves back and inserts the tools into the spindle and magazine and then returns to
its original standby position. The complete tool change takes about 8 sec, providing
the tool search has been completed beforehand.
Tool codes are made up by assembling coding rings and spacers on the tool holder in
a predetermined pattern or code. A combined total of 10 rings and spacers arranged
in a given pattern completes a tool code. This arrangement identifies each toolholder
as it passes a reading head while the tool magazine is rotated. In operation, the
tool-hole pattern in the tape is read, and the machine’s control system calls for the
tool transfer to the machine spindle. The tool magazine then begins to rotate and the
reading head feels the coded ring stack on each toolholder as it passes under the
reading head. When the proper ring combination matches the tape, the magazine
stops. The change is then completed as previously described.
The coding collars on each holder are made up of Group Tool
various combinations of 10 rings and spacers. Half Ring position Ring position
No. 1 2 4 8 16 1 2 4 8 16 No.
the rings identify the group within which the tool
1 1
is coded, while the other half identify the number 2 2
of the tool within the group. Thus, it is possible to 3 3
code the available 31 tool positions within each of 4 4
5 5
31 different groups. Each tool can then be coded
6 6
within each group, making a tool of 961 possible 7 7
tool codes. 8 8
9 9
The chart shown in Fig. 12.75 indicates the number 10 10
of coding rings and spacers required to make up 11 11
any one of the 961 possible tool codes. A shaded 12 12
13 13
square under a ring-position number and opposite 14 14
a group or tool number indicates a coding ring is 15 15
used in that position. Unshaded squares indicate 16 16
a spacer in that position. For example, if tool code 17 17
18 18
0718 is called for, it is established that 07 is the 19 19
group number the 18 is the tool number. In the 20 20
group column and opposite the number 7, shaded 21 21
squares appear only under ring positions 1, 2, and 22 22
23 23
4, since 1 + 2 + 4 = 7. Therefore, there coding 24 24
rings and two spacers are required for the group 25 25
number. In the tool column, opposite 18, shaded 26 26
squares appear only under ring positions 2 and 27 27
28 28
16, since 2 + 16 = 18. Hence, two coding rings 29 29
and three spacers are required for the tool number 30 30
and a total of five coding ring and five spacers are 31 31
needed to make up the complete tool code. Figure
Fig. 12.75 Tool-code chart (The
12.76 shows how the tool code is checked with the
Kearney and Treaker
tool-code chart.
Corporation)
812 Tool Design

Fig. 12.76 Checking tool code with tool-code chart (The Kearney
and Treaker Corporation)

A less sophisticated but efficient and reliable tool changer is shown in Fig. 12.77. Here
the tool magazine has a capacity of 15 toolholders. At the beginning of a program, an
address letter T is used to designate the first tool. The first toolholder has an adapter
Tool Design for Numerically Controlled Machine Tools 813

in the end which identifies it as the first tool. In operation, the T signal is read from the
tape, and the magazine begins to rotate in a clockwise direction until the first tool is in
the transfer position. The tool change is completed and the first machining operation
begins. Thereafter, the magazine indexes one cartridge position for each succeeding
tool change. This means that the toolholders must be placed in counterclockwise
sequence in the order in which they are to be used.

Fig. 12.77 Tool changer on Milwaukee-Matic series Eb (The Kearney


and Trecker Corporation)

Figure 12.78 shows the Edlund-Matic Model 2 NL machining centre, which machine
incorporates an automatic tool changer comprising 15 separate tool-storage pockets.

Fig. 12.78 The Edlund-Matic model 2NL machining centre (Edlund Division of Monarch
Machine Tool Company)
814 Tool Design

Tool diameters up to 75 mm can be changed in less than 5 sec. The tools may be
selected at random. The tools are stored in a vertical position to avoid interference
with the workpiece.
In use, the tools are placed in tool-storage pockets in an inverted position, and each
tool compartment is sequentially numbered, 01 through 15. Any tool may be selected,
manually or by tape command, by programming a T and the two-digit tool-pocket
number or, in the semi-automatic model, selecting the tool through thumbwheel
switches located on the operator’s console.
The tool information being acted upon by the control causes the tool-clamp fingers
to grasp the tool that is presently in the spindle chuck. The clamping action activates
a limit switch, which in turn releases the tool-driving fingers inside the spindle chuck
and thus frees the tool from the chuck. The tool-clamp assembly then swings through
an are of approximately 200° through the clearance between two tool stations in the
storage rack. As the tool passes through the end of its normal arc, the storage rack is
made to rotate in a counterclockwise direction, placing the toolstorage pocket directly
underneath the clamped tool. The tool is deposited into the tool pocket and released.
The tool rack beings to rotate in a clockwise direction until the commanded tool arrives
at the centre of the tool-clamp assembly. The new tool is clamped and lifted from its
pocket. The storage rack moves in a clockwise direction to allow clearance for the
tool to be swung through the space between the tool stations and up into the spindle
chuck. When the tool arrives at the right location for tool-driving fingers to clamp the
tool adapter, the tool-chuck limit switch is activated and the tool is chucked. The tool-
clamp fingers are then retracted out of the way, where they remain until another tool
change is requested (see Fig. 12.79).

Fig. 12.79 The mechanism of the Edlund-Matic 15-position automatic tool changer (Edlund
Division of the Monarch Machine Tool Company)

The Hydrotape machine centre shown in Fig. 12.80 utilises a tool changer that
automatically inserts and secures toolholders in position in the spindle nose by
hydraulic pressure. Figure 12.81 shows the operating principle of the tool changer.
Twelve toolholders are inserted into the tool changer through a door with dotted lines
(A) on the right-hand side of the changers. An index button located on the machine’s
electric cabinet adjacent to the tool-loading door will allow the operator to index, after
Tool Design for Numerically Controlled Machine Tools 815

inserting a toolholder, to the next tool station. The tool-loading door is electrically
interlocked so that when it is opened, all machine functions are cancelled.

Fig. 12.80 The Brown and Sharpe Hydrotape numerically controlled machining centre (The
Brown & Sharpe Mfg. Co.)

Fig. 12.81 The Brown and Sharpe Hydrotape automatic changer


(The Brown & Sharpe Mfg. Co.)

Assuming that all 12 toolholders have been inserted into the tool changer, they are
rotated by means of the spider arms which separate the tools and by a hydraulic
motor. This hydraulic motor will stall if any tool meets an interference while indexing,
with no resulting damage to the tools or fixture.
816 Tool Design

As part of the tool changer there is a tool carrier (B), which advances and retracts
the toolholder. When fully retracted, a cavity in the tool carrier permits the rotating
toolholders to pass through when indexing around. The hydrostatic spindle, which
retracts with the tool carrier, continues to retract until it stops to clear the end of the
toolholder shanks by approximately 1.5 mm as they rotate below. When a tool number
is called for, either manually or by tape, it indexes and stops into position in the tool-
carrier cavity directly below the spindle. A female taper hole on the end of the spindle
contains two spring-loaded drive pins, one 6 mm longer than the other. As the spindle
advances, this longer pin engages a drive slot in the shank of the toolholder, causing
the inner portion of the toolholder to rotate at the same speed as the spindle prior to
the mating of the tapers. The outer face of the toolholder bearing is held securely in
the tool carrier. With this arrangement, it is almost impossible to score up or wear the
tapers in the spindle or on the tool holder shank.
Indexing is random and not sequential. Any tool can be called for on tape using a T01
to T12 command.
Station-to-station indexing time is 3 sec, and indexing 360° takes 7 sec. The chip-
to-chip time averages 6 sec, depending on the spindle extension. This is the time
required to retract a tool, index to the next tool, and start machining the work with
this tool.

12.9 TOOL PRESETTING


Presetting cutting tools involves positioning the cutting edges for their correct
relationship to the workpiece completely away from the parent machine. This allows
the parent machine to stay busy on another job, thus reducing the nonproductive time
required to set the cutting tools by conventional methods.
Tool presetting is usually accomplished through the use of a presetting fixture that
has been designed for a specific machine or a specific system of tooling. Some
presetting fixtures closely resemble the tool positioner of the parent machine itself,
while others are universal in nature. Universal setting fixtures of this type are suitable
for a system of tooling that accommodates several different machines. In some cases,
conventional gauging equipment, such as surface plates, height gauges, indicators,
and optical comparators, any be used. It is the purpose of this section to discuss the
various types of presetting fixtures used and to familiarise the tool designer with the
basic concepts of tool presetting.
Tool presetting is usually done at a tool-presetting centre. This may be the existing
tool crib or a special area of the tool crib designated as a tool-setting centre. The
cutting tools for the N/C machine are stored here, along with work-holding fixtures and
special tooling for N/C machines. The centre is also responsible for the maintenance
of tooling, such as cleaning, inspecting, and replacing cutting inserts.
In operation, when a job is scheduled for an N/C machine, a copy of the work order
is sent to the presetting centre ahead of time. A tooling sheet attached to the work
order contains all the information needed concerning fixtures and cutting tools for the
work to be done. The tool setter (job title for person who presets tools) assembles the
specified tools and adjusts and sets the tools by means of the presetting fixture. The
tools are then loaded on a portable cart. When the preset tools are to be transferred to
the machine according to portable cart. When the preset tools are to be transferred to
Tool Design for Numerically Controlled Machine Tools 817

the machine according to position or station numbers, numbered holes in the cart are
used to keep the tooling in proper sequence. Some companies place tools, fixtures,
tape, and complete job information on the cart for delivery to the machine ahead
of time. Figure 12.82 shows a tool setter adjusting tools by means of a Sundstrand
Optical tool-setting gauge. Note the portable tooling cart. Figure 12.83 shows a typical
pallet cart used to transport heavy tooling.

Fig. 12.82 The Sundstrand optical tool- Fig. 12.83 Typical pallet cart (The Kearney
setting gauge and tooling cart and Trecker Corporation)
(Sundstrand Machine Tool)

Once the new machining operation is ready to being, the machine operator can
easily switch tooling and begin the new job. Additional tool adjustment should not
be necessary except for close-tolerance work. Preset tools should hold open part
tolerances (such as 0.05 to 0.10 mm) without difficulty. Closer tolerances may require
trial cuts and readjustment of tools. The operator should always check the first part to
make sure that the tools have been set correctly.

Presetting End-cutting tools The least complicated method of tool presetting


involves the setting of end-cutting tools, such as drills, end mills, reamers,
countersinks, counterbores, and taps. In this case, it is only necessary to set the end
of the tool longitudinally in relation to a master locating surface. The radial relation
is permanently established by the diameter of the tool. Longitudinal adjustment is
accomplished through an adjustable spindle adapter, as shown in Fig. 12.84, or the
tool itself is moved longitudinally in the spindle adapter. When the tool itself is moved
in the adapter as in the case of collet adapters, the tool has a tendency to shift as the
collet is tightened. Thus, the tool setter must known how much the tool will shift and
make allowances before tightening. Figure 12.85 shown a spindle adapter where the
tool is adjusted within the adapter.
The simplest tool-setting device for longitudinally setting end tools is a surface plate
bored to hold bushings that accommodate the spindle adapter. The bushing represents
818 Tool Design

the machine spindle nose and is used to simulate spindle mounting conditions. A
standard height gauge and dial indicator may be used to set the required gauging
dimension (see Fig. 12.86).

Adjusting nut

20 10
30
Fig. 12.84 Adjustable spindle adapter (Adjusted longitudinally)

Fig. 12.85 Collet-type spindle adapter. Cutting tool is adjusted longitudinally within adapter
before tightening.

Fig. 12.86 Tool presetting with a standard height gauge


Tool Design for Numerically Controlled Machine Tools 819

Commercial presetting fixtures that operate on the same principle are available.
Figure 12.87 shows a presetting height gauge in which a master adapter holder is
mounted vertically n the base. The base also contains a vertical arm containing stop
spaces 25 mm apart. A removable forked arm is provided for setting various tool
lengths. Other models have a 25 mm micrometer head attached to the forked arm
to allow adjustment within ± 0.025 mm. The base is drilled and counterbored for
mounting on a bench, allowing it to be used as a locking fixture when collet-type
adapters are used.
A similar tool setter shown in Fig. 12.88, is known as the Econumeric and is
manufactured for use in presetting tools for the Milwaukee-Matic machining centres.
The tool adapters are inserted into a precision-bored hole and locked in place by a
sliding fork. Longitudinal lengths are established by a surface gauge or height gauge.
Fine adjustment to the tool is made by a small handwheel attached to a pushrod that
extends through the bottom of the tool adapter. Collet-type adapters may be locked
directly in the spindle.

Fig. 12.88 The Kearney and Trecker Econ-


Fig. 12.87 Universal Kwik-Switch preset- umeric tool presetting fixture
ting height gauge (The Universal (The Kearney and Trecker Cor-
Engineering Company) poration)

In addition to longitudinal settings, the Econumeric tool setter will check radial setting.
The tool adapter is held between small rollers, thus allowing the adapter and cutter
to rotate. The distance between the centreline of the tool and the reference surface
of the tool setter is known, thus allowing radial adjustments to be made with surface
gauge or height gauge and dial indicator. This arrangement is also quite useful in
checking cutter diameter, tooth height, cutter relief, and radial runout.

Presetting radial-cutting tools A presetting spindle designed specifically for the


radial setting of boring bars is shown in Fig. 12.89. It was developed to provide a
method of accurately presetting boring bars to the required bore diameters prior to
their use in the machine-tool spindle. The standard model is equipped with no. 40
and no. 50 combination Flash Change and National Machine Tool Builders taper
sockets in a double-end design that allows tooling to be preset for a variety of machine
tool spindles.
820 Tool Design

Fig. 12.89 Precision presetting spindle (DeVlieg Microbore, Division of


DeVlieg Machine Company)

In use, the precision presetting spindle is mounted on a surface plate. The gauge
reference surfaces of the spindle housing represents the centreline of the spindle.
Gauge blocks equaling the required radius dimension are positioned on the gauge
reference surface, and an indicator is adjusted to zero to establish the tool-setting
dimensions (see Fig. 12.90). The cutting tool is then adjusted to the zero indicator
setting and the spindle rotated manually to assure that the setting is made to the high
point, or crest of the tool-point radius.

Fig. 12.90 Use of precision preset spindle (DeVlieg Microbore, Division of


DeVlieg Machine Company)

A more expanded version of the presetting spindle (Fig. 12.91) is known as the DeVlieg
Precision Presetting Machine. It will preset boring bars and cutting tools to accurate
depth dimensions as well as accurate diameter. It consists of a precision presetting
spindle mounted on a cast-iron base and is arranged with transverse and longitudinal
slide movements. Precision optical scales that can be read directly on 0.025 mm
without interpolation are provided to establish diameter and length measurements.
Tool-point gauging is accomplished through the optical-comparator principle using a
40-power microscope with a precision cross-hair reticle.
In use, a gauge bar is positioned in the presetting spindle, as shown in Fig. 12.92,
and the optical scales for depth and diameter measurements are positioned to zero
reference points. The two knife-edge surfaces on the gauge bar permit clear alignment
with the cross-hair reticle, and the longitudinal scale for depth setting is zeroed to the
Tool Design for Numerically Controlled Machine Tools 821

Fig. 12.91 DeVlieg precision preset- Fig. 12.92 Zeroing scales of precision
ting machine (DeVlieg presetting machine (DeVlieg
Microbore, Division of Microbore, Division of DeVlieg
the DeVlieg Machine Co.) Machine Company)

gauge line of the taper-spindle socket.


The transverse scale for diameter
setting is zeroed to the centreline of the
spindle.
The gauge bar is removed, and tooling
with axial and radial adjustments is
inserted into the spindle, as shown
in Fig. 12.93. The transverse slide is
adjusted until the optical measuring
element reads out the required radius
dimension. The diameter is then set
by adjusting the boring unit until the
high point, or crest of the tool point,
is positioned (while rotating the bar) Fig. 12.93 Presetting tools or precision pre-
to the horizontal cross hair of the setting machine (DeVlieg Micro-
microscope reticle. bore, Division of DeVlieg machine
The longitudinal slide is adjusted until Company)
the optical measuring unit reads out the
depth setting required. The boring bar is adjusted through its axial adjustment until
the lead cutting edge of the boring tool is positioned to the vertical cross hair of the
microscope reticle.
Presetting machines comparable to the DeVlieg unit described above are shown in
Figs 12.94 and 12.95. Figure 12.94 shows the floor-mounted Kearney and Trecker
Optical Tool Setter. It is designed to preset tools for Milwaukee-Matics and other
machines that use adjustable toolholders. Like the DeVlieg unit, it contains a
single monocular tool-setting microscope and optical measuring scales and reads
with minimum measuring increments of 0.0025 mm in either the horizontal or
vertical direction.
822 Tool Design

Fig. 12.94 The Kearney and Trecker optical Fig. 12.95 The Sundstrand optical tool-
tool setter (The Kearney and setting gauge (Sundstrand
Trecker Corporation.) Machine Tool)

Figure 12.95 shows the Sundstrand Optical tool-setting gauge. It is a high-precision


measuring instrument for accurately presetting Sundstrand Omnitools as well as tools
for other preset tooling systems. Optical viewing of the tool by means of a 42-power
microscope assures accurate measurement and at the same time allows inspection
of the tool tip. Digital optical measuring instruments on each axis provide fast,
reliable, and direct readout of the slide position to 0.0025 mm. The toolholding fixture
incorporates a Hydra-Lock chuck. A second station is available for presetting of 37
and 56 mm straight-shank tools, DeVlieg Flash Change tools. Lathe tools, and tools
incorporating the no. 50 standard milling-machine taper.
A method of presetting tools that has not been mentioned thus far is the use of a
standard optical comparator. Many manufacturing companies have optical comparators
in their inspection departments which have enough capacity for staging cutting tools
and their adapters. All that is necessary to convert it to a tool presetter is to design
a stage suitable for holding the tool adapter. The stage could be a V block, a master
spindle adapter, or a tool post for a quick-change tool system (see Fig. 12.96). Once
the tool adapter is mounted, measurements can be taken by conventional optical-
comparator method.

Presetting tools for N/C lathes Many companies believe that it is more practical
to preset tooling for N/C lathes than for N/C machines employing rotating cutters.
This is due to the incorporation of tool offsets in the control design. Tool offsets allow
the operator to correct the position of the tool for errors in the cross and longitudinal
Tool Design for Numerically Controlled Machine Tools 823

direction. Thus, tools may be preset for one


par tolerances (0.05 to 0.1 mm) and handled
in ordinary fashion, but closer-tolerance work
requires taking a finish cut, “miking” the part,
and dialing in the correction by means of tool
offsets supplied with the control.
A pair of offset controls is required to shift the
lathe tool in the cross (x) direction and in the
longitudinal (z) direction. When the tool is held
in a turret, each pair of offsets may be assigned
to any tool in the turret by tape code. When an
offset is assigned to a turret station (station 1, for
example), the tool’s position is actually moved
in the x and z direction by the amount dialed in
at the controls. When a new tool is called for
(no. 4 for example), the old offset magnitude is
removed an the new offset magnitude dialed
in for tool no. 4 causes the tool to the proper Fig. 12.96 Utilising the standard
location. With this approach, high accuracy is optical comparator as a
achieved with a minimum of spindle idle time. tool-presetting fixture
The open tolerances that require no offset (The Jones & Lamson
compensation cost essentially no spindle idle Machine Tool Company)
time.
Several manufacturers have incorporated optical comparators in the construction
of tool-presetting machines for N/C lathes. Typical examples are shown in the
following pages.
Figure 12.97 shows a presetting
machine designed for presetting
tools on N/C lathes manufactured
by the Lodge and Shipley Company,
known as the Lodge and Shipley
Numeriset. The comparator lens is
mounted in a vertical plane directly
above the Numeriset centreline. The
toolholder adapter is located on a
horizontal plane, with its reference
surfaces corresponding precisely
to the optical-comparator crosslines
when the adapter is in its zero-zero
position. Directly calibrated to the
adapter reference surfaces are optical
viewers and scales which provide
direct readings of the distance from Fig. 12.97 The Lodge and Shipley Numeriset
the adapter x and z surface to the tool-presetting machine (The Lodge
optical-comparator crosslines. The & Shipley Company)
x and z axis distances are directly
related to the lathe, which has movement in these two axes.
824 Tool Design

In use, the tool slides are set by the optical scales to the programmed dimensions. The
tool and holder are then mounted, and any variance will be noted on the comparator
screen. The tool is then shifted until the tool edge and optical cross hairs match.
A similar example is shown in Fig. 12.98. This tool-setting gauge is manufactured for
use with Tape-Turn numerically controlled lathes. It utilizes a Stocker-Yale optical scope
on a bench-type fixture with compound adjustment for x and y axis in increments of
0.0025 mm. It permits the use of standard throwaway insert toolholders and provides
cutting-edge positioning to equal 0.025 in the turret of the lathe.

(a) (b)

Fig. 12.98 The LeBlond Opti-Set: (a) Using master setting gauge (b) Tool in position (LeB-
lond Machine Tool)

In use, the comparator crosslines are


moved by the x-and y-axis slides until they
touch the shadow of the master setting
gauge (see Fig. 12.98a). This establishes
the zero-zero position. A standard insert-
type toolholder is then mounted in the
tool block, as shown in Fig. 12.98b. The
toolholder is adjusted by means of screw-
type locators in the side and end of the
tool shank until the tool tip touches the
comparator centreline.
Any offset form the zero-zero position
can be made by moving the comparator
crosslines the desired distance by the
graduated handwheels on the x and y axis
slides. The tool is then adjusted to comply
Fig. 12.99 The Monarch tool-set-
with the desired amount of offset.
ting gauge for use with the
The Monarch Machine Tool company smaller Monarch Pathfinder
offers two models of tool-setting gauges numerically controlled lathes
for use with their numerically controlled (The Monarch Machine Tool
Monarch Pathfinder lathes. The main Company)
Tool Design for Numerically Controlled Machine Tools 825

difference between the two is that one provides for direct installation of the lathe turret,
while the other allows the tools themselves to be preset prior to their introduction to
the lathe toolholders. The reason is that as the lathes get larger, it is more difficult to
interchange turrets because of turret size and weight and still retain good efficiency.
Figure 12.99 shows the Monarch tool-setting gauge for the smaller Monarch Pathfinder
lathes. It consists of a base, a carriage to support the cross and upper slides, and
an indexing post. Note the turret and tools mounted on the indexing post. The post
will accept either square or hex turrets. The tools are brought into approximate
position by rotating the indexing post manually. Final and accurate turret positioning is
accomplished by raising the turret position-locating lever.
The upper slide (on which the optical-comparator gauging device is mounted) is used
to extend the range of tool offset that can be gauged. It is used in only two positions:
Full out and clamped against the stop or full in and clamped against the stop.
The unit is calibrated by means of a master gauge. The slides are moved until the
cross hair is properly located on the master gauge. The Trav-A-Dials (a form of dial
indicator) are zeroed, and the carriage is then retracted to allow the turret to be
mounted on the indexing post.
Once the turret is installed, the tools can be mounted and positioned in their proper
stations. The turret is rotated to move the desired tool into gauging position, and the
turret position-locating lever is raised to finish turret positioning. The carriage and
cross slide are then moved to the dimensions. Indicated on the tooling drawing by
utilising the Trav-A-Dials. The tool is adjusted manually with in the turret station until
the tool cutting edges are properly located in the optical-comparator cross hairs.
Height is established by tool-shank and turret-surface accuracy or by use of shims to
establish the desired height accuracy. The carriage is retracted to allow the next tool
to be rotated into gauging position.
The Monarch tool-setting gauge used for setting tools for the larger Pathfinder lathes
allows the tools themselves to be preset prior to their introduction to the toolholders on
the lathe (see Fig. 12.100). The holders are accurately machined to accept the preset
tools with precision. The tools are supplied with calibrated adjusting screws located
in the tool shank, as shown in Fig. 12.101. The length (x-axis) adjustment is made by
means of a shoulder fixture mounted on the tool, while the side (z-axis) adjustment is
made by means of two calibrating screws located on the side of the tool shank.
The gauge itself s similar to the Monarch gauge previously described, except that a
tool support is used in place of an indexing post. The tool support provides a receiver
for both right- and left-hand tools as well as one for boring bars and drills. The full
range of tools for all Monarch Pathfinder lathes is accommodated by using only two
tool-adapter blocks. These are factory-calibrated and can be installed in seconds.
Using various-diameter boring bars is facilitated by means of toolholder bushings.
Calibration and operation are similar to those descried for the previous model. The
accuracy to which both models of the tool-locating fixture can be positioned is said to
be 0.025 mm in 300 mm.
826 Tool Design

Fig. 12.100 The Monarch tool-setting gauge for use with larger Monarch Pathfinder numeri-
cally controlled lathes (The Monarch Machine Tool Company)

x-axis adjusting
screw

z-axis adjusting
screws

Fig. 12.101 Lathe tool in hex turret showing adjusting screws (The Monarch
Machine Tool Company)

Summary

The conventional machine tools are being replaced by numeric control machine
tools. Numeric control is a significant step toward job-shop automation. Numerically
controlled machine tools are capable of much superior accuracy, precision and finish,
compared to conventional machine tools. Numeric control reduces the set-up time,
time required for change of speed and feed, time required for tool change, and time
required of workpiece indexing. There is a need of precise tooling and fixturing for
such numerically controlled machine tools. This chapter discusses the tooling and
fixturing aspects of numerically controlled machine tools.
• Strap clamp and heel block build in to one unit is well suited to fixturing for N/C
machine tools.
• Work supports and locators may also be incorporated into one unit for use in
N/C machines.
Tool Design for Numerically Controlled Machine Tools 827

• An auxiliary base plate containing grid pattern of holes is especially suited for
N/C drilling machines.
• Modular fixturing system in often incorporated for N/C machining centres.
• The use of N/C machines has simplified or eliminated elaborate jigs or fixtures
but it has increased the need for exact and rigid cutting tools.
• Drill guide bushings are not required in N/C drilling machines.
• N/C drilling machines requires a drill with drill point that will centre itself without
walking. This is achieved using spiral drill point.
• N/C milling machines are programmed according to specific cutter diameter.
Reduction in cutter diameter due to wear will cause inaccuracy. Therefore, it is
preferable to use insert-type milling cutters.
• N/C machines are equipped with quick changing tool holders, automatic tool
changers and positioners.

Questions
1. What led to the development of numerical control?
2. What are the two basic systems of numerical control in use today?
3. Why is a computer generally used when programming continuous-path N/C
systems?
4. What is a N/C machining centre?
5. Why can machining time on N/C machining centre be predicted?
6. Generally speaking, what is the ultimate goal when designing fixtures for N/C
machine tools?
7. What modifications are made to a standard milling-machine vise to convert it
to a N/C vise?
8. How are lathe chucks used as holding fixtures in N/C machine tools?
9. What is the advantage of universal gird grid plate for N/C tooling?
10. What is meant by the Tinker Toy or Erector set fixturing concept?
11. Why is the standard twist drill usually not suited for precision drilling operations
on a N/C drilling machine?
12. What is the first requirement of a drill to be used on a N/C machine?
13. How is it possible to establish drill location when using standard drills in a N/C
drilling machine?
14. What is the correct depth of a spotting operation to locate a hole pattern?
15. How is it possible to get shorter flutes using standard twist drills?
16. Why should collet-type drill holders be used in N/C drilling in place of standard
drill chucks?
17. What is the advantage of spade drills?
18. Why do N/C milling machines require heavier and stiffer milling cutters?
19. List the various ways of overcoming the problem of undersize milling cutters
that have been reground.
20. Why is tapping on N/C machines generally easier than tapping on conventional
machines?
828 Tool Design

21. What is the best type of tap to use on through holes when tapping on a N/C
machine? Why?
22. When is spiral-fluted tap used? Why?
23. What type of tap holder is used on N/C machines that do not have provisions
for adjusting the feed rate to accurately match the lead of the tap?
24. Why can single-point throwaway insert tooling be considered quick-charge
tooling?
25. How does the operator know when to change tools on N/C machines that
require manual change of tools?
26. What is unique about the Erickson system of quick-change tooling?
27. What is the advantage of an automatic tool charger when it is sometimes pos-
sible for a good operator to change tools faster manually?
28. What is the tool-length compensator built into the controls of turret-type N/C
machines?
29. How does the Kearney and Trecker automatic tool charger identify the tool
needed for a particular machining operation?
30. Referring to Question 29, how are tool codes made?
31. What is the maximum number of possible tool codes for the Kearney and
Trecker tool changer?
32. How is tool charge called for on the Edlund-Matic machining centre?
33. What is meant by tool presetting?
34. Where is the tool presetting actually done?
35. How close can tolerances be held with preset tools?
36. What is the most common and least complicated method of tool presetting?
37. What modifications are necessary to convert a standard optical comparator
into an optical tool setter?
38. What is meant by a tool offset that is built into the control of a N/C lathe?

Design Problems
1. Select a N/C machine tool and design a system of universal tooling that will
allow the set-up of a wide variety of workpieces. Use an existing N/C machine
tool or select a N/C machine tool from the manufacturer’s brochure for measre-
ments and dimensions.
2. Make a tool drawing showing the alterations in drill-point geometry on a stan-
dard twist drill to be used on a N/C drilling machine.
3. Select a N/C machine tool and design a tool-setting fixture to present the tools
for that particular machine. Use an existing N/C machine tool or select a N/C
machine tool from the manufacturer’s brochure for dimensions and tooling
systems.
4. Select a N/C machine tool and design a tooling cart to transport and carry tool-
ing from the tool-setting centre to the individual mahcine.
5. Design a universal N/C fixture to hold the workpiece shown in Figs. 8.70, 8.67a,
and 8.67b.
Tool Design for Numerically Controlled Machine Tools 829

References
Books
• Childs, JJ: “Principle of Numerical Control,” The Industrial Press, New York, 1965. IIT, “APT Part
Programming”, McGraw-Hill, New York, 1967.
• Robert, A D and R C Prentice: “Programming for Numerical Control Machines,” McGraw-Hill,
New York, 1968.
• Thornhill, RB: “Engineering Graphics and Numerical Control”, McGraw-Hill Book Company,
New York, 1967.
• Wilson, FW (ed): “Numerical Control in Manufacturing”, McGraw-Hill, New York, 1963.
Catalogues
• Aloris Tool Co., Inc., Clifton, NJ.
• Avey Machine Tool Company, a Division of the Motch & Merryweather Machinery Co.,
Clincinnati, Ohio.
• Burgmaster Division of Houdaille Industries, Los Angeles.
• Cincinnati Milavron, Cincinnati, Ohio.
• Colt Industries, Inc., Partt & Whitney, Inc., West Hartford, Conn.
• DeVlieg Machine Tools, Royal Oak, Mich.
• Edlund Division, Monarch Machine Tool Company, Cortland, NY.
• Erickson Tool Company, Solon, Ohio.
• Jones and Lamson, Springfield, Vt.
• Kearney and Trecker Corporation, Milwaukee, Wis.
• LeBlond and Shipley, Cincinnati, Ohio.
• Lodge and Shipley, Cincinnati, Ohio.
• The Monarch Machine Tool Company, Sidney, Ohio.
• Moore Special Tool Company, Inc., Bridegeport, Conn.
• Sundstrand Machine Tool, Belvidere, Ill.
• Universal Engineering Company, Frankenmuth, Mich.
13 AUTOMATIC
SCREW MACHINE

13.1 INTRODUCTION
The automatic screw machine was developed from lathe for the economical
manufacture of small screws and bolts. Today this machine is by no mean limited
to screws but is capable of producing a wide variety of products. Important parts
produced wholly or partially or partially on automatic screw machines include shafts,
compressor nuts, threaded retainer rings, knobs, tapered plugs, spark-plug shells,
and artillery shells.
The principal features of the automatic screw machine are automatic stock feeding
and operation of all cutting tools by cam action. Once the machine is set-up, supplied
with stock, and started, it needs little or no attention from the operator. In this way,
one operator can look after several machines, thus ensuring economical labour costs.
Further economy results from the fact that changes in cutting-tool position and stock
feeding are made much faster automatically than by hand.
There are many types of automatic screw machines. Some have single spindles, while
others have several spindles arranged within an indexing head. Some can handle only
bar stock, while others, known as chuckers, can machine forgings or castings with the
aid of hand or magazine feeds. All carry cutting tools in turrets and in cross slides.
Furthermore, all are built to perform several cutting operations at once or in as rapid
a sequence as possible. The actual cycle of operation is carefully planned to reduce
production time to a minimum and maintain the required quality of work.
Since the layout and design of the special cams that control the cutting tools of
automatic screw machines is a specialised area of tooling, the average tool designer
may never be confronted with this problem. Companies that specialise in screw-
machine products may do their own cam design, but even they are finding that it
is often more economical to secure the services of companies specialising in cam
design and production. However, the tool designer should understand the principles of
automatic screw machines, and for this reason, a description of one type of machine
is presented here.

13.2 GENERAL EXPLANATION OF THE BROWN


AND SHARPE MACHINE
A typical Brown and Sharpe automatic screw machine is shown in Fig. 13.1, along
with operating controls and principal parts. It will readily be seen that the machine
is a specially designed lathe equipped with cams, levers, clutches, and trip dogs
for actuating and controlling the cutting tools without the aid of the operator once
the machine is set up for a particular job since the stock is always held in collets or
chucks. No tailstock is used, but a six-hole turret is mounted in its place. The stock is
automatically advanced the right amount after the finished part has been cut-off until
Automatic Screw Machine 831

14 13 12 11 10 9

15
1

16 2

17
6
18
19
3
20 8 7
4

22
21
5
23

24

26

25

Fig. 13.1 Operating controls and principal parts of Brown and Sharpe Ultramatic Screw Machines
(Brown & Sharpe Mfg. Co.)
1. Eye-level control station has spindle start and stop buttons and driveshaft start, stop, and jog buttons.
Provision is also made for mounting on-off switches for attachments. 2. Turret and speed –change dog
carriers are directly below control station for convenience in set-up, with lead cam visible. Graduation
on each carrier to hundredths of a revolution simplify both the layout of a new job and duplication of
previous setups. 3. Cycle change gears are readily accessible in the compartment at the right end of the
machine. Four gears on one fixed and one adjustable centre are retained by the electrically interlocked
safety cover. 4. Handwheel for driveshaft rotation is located at the right front of the machine. It oper-
ates at 3:1 ratio for easy manipulation. A positive interlock ensures disengagement when machine is in
operation. 5. Lead cam for the turret slide is mounted on the camshaft at the lower right, protected by
Plexiglas shield. The turret slide can be operated manually by an easily applied lever. 6. Turret slide has
constant stroke and maximum positive pull-off of 59.58 mm for each index and rapid pullout of 63.55 mm
for deep hole drilling. 7. Six-position turret has heat-treated shank mounted on preloaded, tapered roller
bearings. (10-hole turret, same diameter, available at extra cost.) 8. Swing stop for stock permits using
all turret holes for tools or using single turret tool without indexing. Has fine adjustment for longitudinal
positioning. 9. Red light indicates when driveshaft is stopped and stock tube empty. 10. Adjustable work
lamp is standard equipment. 11. Third slide (and optional fourth slide) use 177.94 or 203.36 mm cams,
either solid or slotted. Most cams from older model machines can be used. 12. Electrical-control cabinet,
at top rear of machine, is readily accessible. 13. Sliding transparent splash guard permits easy access to
working areas. 14. Counter for positive stopping mechanism is set to stop driveshaft and drive sleeve when
the last piece in the bar is completed. 15. Spindle direction selector (not visible in this view) is located
on the end of the control cabinet. It controls the direction of spindle rotation in high speed. 16. Spindle
(under hinged guard) is precision-bearing, reversing, removable-unit type. It has 5.595 KW chain drive at
all speeds, from independent motor. 17. Spindle –speed table easily visible at eye level. 18. Lubricating-oil
pressure gauge provides visual monitor of system’s operation. 19. Multistation control drum simplifies
settings for programming auxiliary machine functions. 20. Duplicate jog button is provided next to the
multistation control drum at the left front of the machine (not visible in this view). From any operator
position, there is a jog button within easy reach. 21. Oil gauge and filler opening for reservoir supplying
the mechanical lubricator. 22. Front camshaft operates under slides, chucking and feeding mechanism,
and the cross slides. 23. Spindle-speed change-gear compartment. Gear storage in compartment door.
Base lubrication reservoir and oil-level gauge behind door. 24. Speed case and access cover. 25. Feed
change-gear storage behind louvered cover. 26. Drain plug aids in quick coolant-oil changing.
832 Tool Design

the bar is used up. Bars are 3, 3.65 or 6 m in length, depending on the floor space
available, the most widely used being the 3 m bar.
The following sections explain in detail the purpose and use of the controls of the
no. 2 and 3 Brown and Sharpe Ultramatic Screw Machine.

Drive The spindle of the machine shown in Fig. 13.1 is driven by a 5.595 KW motor,
which is a constant-speed type mounted on an adjustable bracket inside the base
(see Fig. 13.2). Drive to the spindle is transmitted through a chain and sprocket to
speed and ratio change gears in the speed case and thence by chains and sprockets
to the spindle.

Fig. 13.2 Spindle driving motor and speed case in base of machine, guards removed (Brown
& Sharpe Mfg. Co.)

The driveshaft, chuck, and feed cam drive


at the rear of the machine are powered by
a 373 W friction-brake motor. A plate on
the back guard indicates the placement
of gears to give either 120 or 240 rpm of
the chuck and feed cam drive, with the
spindle running in either direction (see
Fig. 13.3). The speed of the driveshaft is
a constant 120 rpm with the driving motor
fully loaded.

Spindle The machine is equipped with


one of four spindles- 19.06, 31.77, 41.30,
or 50.84 mm capacity (60.37 mm capacity
with outside feed and heavy-duty stock
loader). All are of unit construction with Fig. 13.3 Speed change gears for
antifriction bearings and easily removed driveshaft, chuck, and feed
from the machine when necessary. These cam (Brown & Sharpe
spindles are mounted in boxes supported Mfg. Co.)
in the bed of the machine.

Spindle speeds Two-speed machines Spindle speeds vary with the capacity of the
spindle. Table 13.1 indicates the speeds that are available with each size.
Automatic Screw Machine 833

Table 13.1 Spindle speeds of two-speed ultramatic screw


Machine Speed range, No. of high High-speed Second Ratio of high
rpm speed range, rpm speeds* to second speed
No. Capacity mm

2 19.06 5018-20 18 5018-34 15 1.6:1-15.3:1


2 31.77 3503-14 18 3503-219 15 1.6:1-15.3:1
2 41.30 2906-14 17 2906-219 15 1.6:1-15.3:1
3 50.84 2484-14 17 2484-187 15 1.4:1-13.1:1
and 60.37
* Available for each high speed.
SOURCE: Brown & Sharpe Mfg. Co.

High and low speeds can be either


forward or backward, but it is advisable
to run both spindle sprockets in the
same direction when the spindle is used
constantly in one direction.
6
Change gears for varying the spindle
speeds are located in the left end of the
base, with gears not in use stored in 5
the compartment door (see Fig. 13.4). 4
The gears are easily removed from the
shafts by loosening the clamp screws to Clamp
slide off the slotted retaining washers. screw and
3 slotted
For gear installation, the procedure is washer
reversed. 1 2

Four-speed machines As with two-


speed machines, the spindle speeds
vary with the capacity of the spindle.
Fig. 13.4 Spindle-speed change gears
Table 13.2 indicates the speeds available
and base oil reservoir (Brown
with each spindle size. In these machines,
& Sharpe Mfg. Co.)
two of the speed changes are made with
the spindle clutch, and two more change

Table 13.2 Spindle speeds of four-speed ultramatic screw machines


Speed No. of High- Second Ratios Third Ratios of Fourt Ratios
No. Capacity range, high speed spreeds* of high speeds* high to speeds* of high
mm rpm speed range to second third to fourth
rpm speed spreeds* speeds

2 19.06 5018-4 18 5018-314 10 1.2:1-5.1:1 15 1.6:1-15.3:1 150 1.9:1-78:1


2 31.77 3503-3 18 3503-219 10 1.2:1-5.1:1 15 1.6:1-15.1:1 150 1.9:1-78:1
2 41.30 2906-3 17 2906-219 10 1.2:1-5.1:1 15 1.6:1-15.1:1 150 1.9:1-78:1
3 50.84 2484-3 17 2484-187 10 1.2:1-5.1:1 15 1.4:1-13.1:1 150 1.6:1-66:1
and 60.37

*Available for each high speed.


SOURCE: Brown and Sharpe Mfg. Co.
834 Tool Design

are provided by the two multiplate electrically operated disk clutches in the speed
case.

Feed change gears Figure 13.5 shows the feed change gears that govern the cycle
time. They are located at the right-hand end of the machine. Engaging the driveshaft
handwheel or opening the cover automatically stops the driveshaft.

Fig. 13.5 Feed change gears installed at right end of machine (Brown & Sharpe
Mfg. Co.)

The feed change gears provide 448 rates of production (74 are shown on the chart in
Fig. 13.7), from 1.6 to 1,000 sec per cycle of the camshafts. Gears not in use are kept
in the gear compartment at the lower end of the machine as shown in Fig. 13.6. The
feed change-gear plate is located on the change-gear door (see Fig. 13.7). The four
gears used are determined by the time in seconds required to make one piece. The
driver gear is stud is provided to give the proper gear meshing. A nut for clamping the
stud in position is located on the outside rear of the changer-gear bracket.

Fig. 13.6 Feed change gears stored in compartment at lower front of machine (Brown &
Sharpe Mfg. Co.)
Automatic Screw Machine 835

Fig. 13.7 Feed change gear plate (Brown & Sharpe Mfg. Co.)

The spring collet and stock-feed mechanism To feed the stock, the bar is gripped
firmly in the spindle by a spring collet located in a sleeve at the extreme front of the
spindle so that the stock can be held at a point as close as possible to the tools. The
shoulder of the collet bears against the ground inner surface of the chuck nut, which
is screwed up tight against the nose of the spindle.
Stock having a variation of not more than 0.12 to 0.15 mm can be chucked satisfactorily
at one setting of the spring collet. Different collets are used for each size of stock.
Feed-tube bushings as near to stock size as possible are used in the outer end of the
feed tube to steady and support the bar when smaller-diameter stock is being used.
The chuck-clutch trip lever, as shown in Fig. 13.8, controls the chuck clutch that
operates these mechanisms through gearing to the chuck camshaft. The chuck and
feed cam on this shaft operates the chuck fork, which causes the collet to open, stock
to be advanced, and collet to close in correct sequence.
The collet is closed when the chuck fork sleeve (see Fig. 13.9) is at its left position,
and its operation is as follows. When sleeve L slides to the left, the two chuck levers
N whose fulcrums are at M force chuck sleeve L (by means of chuck closing tube)
forward upon the taper of the spring collet and close it firmly upon the bar. On releasing
the chuck levers through the movement of the sleeve to the right, the spring and taper
of the collet are sufficient to slide the chuck sleeve back, and the collet opens. The
bore of the spindle nose centres the sleeve and collet.
836 Tool Design

Fig. 13.8 Cam shafts, dog carriers, cams and trip levers control timing of all operations
(Guards removed) (Brown & Sharpe Mfg. Co.)

The stock is advanced automatically by an adjustable feed slide, as shown in Fig.


13.10. Opening the chuck before stock advancement and the stock feeding cycle are
controlled by adjustable dogs on dog carriers. The chuck is automatically closed after
feeding. In proper operation, the collet takes hold of the stock a few thousandths of an
inch before the feed slide completes its advance.
The time required to feed the stock is 0.5 sec if the chuck and feed cam drive sleeve
(which runs independently of the driveshaft for other machine functions) is operating
at 240 rpm or 1 sec at 120 rpm. With the drive sleeve running at 240 rpm the maximum
feed-slide movement is 60.37 mm (19.06 mm-capacity spindle), 63.55 mm (31.77 mm
and 41.30 mm-capacity spindle), and 57.19 mm (50.84 mm-capacity spindle), or if
running at 120 rpm, 101.68 mm for all four spindles. The feed slide is protected by two
1.58 mm shearing pins located in the plunger sleeves.

The automatic operating mechanism Two driving shafts at the rear of the machine
connected by a chain coupling, are driven at 120 rpm regardless of spindle speed
(see Fig. 13.11).Clutches on the driveshaft control operational movements of the
machine such as indexing the turret and changing the spindle speed.
A third driving unit, which also includes a clutch, rotates over one end of the driveshaft
and drives the chuck and feed cams. This unit can turn at either 240 or 120 rpm,
independent of the driveshaft spindle speed.
The driving clutches, which are fastened to the driving shafts, have teeth on the ends
to engage the sliding clutch members. Inside the body of the sliding members is a
compressed coil spring, which is further compressed when the clutch is disengaged.
When the trip lever is raised by a dog on the carrier, the back end of the lever is
depressed, allowing the sliding clutch to engage as shown in Fig. 13.12. The
spindle-reversing clutch is disengaged after 0.5 rev (or 0.75 rev when using neutral
Automatic Screw Machine 837

Fig. 13.9 Sections through spindles of ultramatic screw machines: (a) 19.06 mm capacity
(b) 31.77 mm capacity (c) 41.30 mm capacity (Brown & Sharpe Mfg. Co.)

clutch pockets). The chuck clutch makes two revolutions and the turret clutch one.
Simultaneously with the disengagement of the clutch, the spring is further compressed,
ready to operate the clutch again.

Camshafts There are two camshafts on these machines (Fig. 13.8). The one located
on the left front of the machine carries (from left to right) the control drum, the upper-
slide cams, the chuck dog carrier, and the cross slide cams. Provision is made, at the
right hand end of this shaft, for coupling to the transfer mechanism or the Longitudinal
Turning Slide camshaft if the machine is so equipped.
838 Tool Design

Shear pin levers


for upper slides Upper slide rear.
swivel clamp screw

Feed slide
adjusting knobe
Upper slide front,
swivel clamp screw

Chuck adjusting Nut

Feed
latch
Chuck fork

Feed scale Spindle clutch fork


and pointer
Spindle rear box

Link adjusting screws


Upper slide
front, lever Upper slide rear, lever

Fig. 13.10 Spindle and feed adjustments, and upper slide adjustments (Brown & Sharpe
Mfg. Co.)

Drive shaft chain coupling


Trip levers (Behind this relief valve)
Pull-out
cam lever Spindle-Reversing
clutch

Camac
mechanism

Lead camshaft
Lubricator and
Trip-lever coolant pump
abutment motor

Fig. 13.11 Rear view of machine (Guards removed) (Brown & Sharpe Mfg. Co.)

The second camshaft runs at right angles to the front camshaft and is located at the
right end of the machine. This shaft carries the lead cam (for turret slide) on its left
end, and the rapid pull-out cam on the right. A “dummy” disc cam replaces this second
cam when the Rapid Pull-Out Attachment is not used.
Automatic Screw Machine 839

Trip lever Clutch


Driving
Trip
shaft
dog

Trip lever fulcrum


Cam shaft
Dog carrier

Fig. 13.12 Trip lever and dog carrier (Brown & Sharpe Mfg. Co.)

Both camshafts are driven from the driveshaft at the rear of the machine, through the
feed change gear, through a non-adjustable mechanical safety clutch, and then by
individual gearing and separate worms and adjustable wormwheels.
When changing a cross slide cam or the lead cam, simply remove the adjacent nut and
slip off the collar or adjustable can holders with the cam. All cams with the exception
of the turret lead cam have adjustable cam holders. The cams are positioned on the
adjustable holders by means of locating pins in the holders and holes in the cams.
These holder are graduated in hundredths of a revolution and provide for adjustment in
the event that cams designed for another job are used. Also, they permit compensating
for error in cam design or manufacture.
The upper-slide cams may be solid or slotted. Removal of these cams requires taking
out bolt A as shown in Fig. 13.13 and moving the inside sleeve as far to the left as it
will go by pulling on the knurled diameter provided. This disengages the clutch driving
the control drum and thus provides space for slipping off the upper-slide cams and
their holders.

Fig. 13.13 Control drum at end of camshaft, left (Brown & Sharpe Mfg. Co.)

Pin levers facilitate manual operation of cross-slide cam levers when installing cams.
A removable lever also simplifies removal of the turret-slide cam.

Dog-shaft carrier This shaft is located above and behind the turret slide. At the left
end of the shaft is the turret dog carriers; at the front is the reversing dog carrier. The
turret dog carrier control the timing of the turret indexes. The reversing dog carrier
controls the spindle clutch and thus controls the spindle-speed change.
840 Tool Design

Trip dogs are accommodated on both sides of the turret dog carrier to allow for close
spacing of dogs when closely spaced turret indexes are required. Normally, half the
dogs are mounted on one side and half on the other.
The dog carriers, including the chuck dog carrier (adjacent to the control drum at the
left front of the machine) are graduated in hundredths of a revolution. This simplifies
setting up a new job and duplication of previous setups.
All dog carriers are driven from a common motor and remain synchronised at all
times.

Trip dogs Dogs on a carrier lift a trip lever that releases a jaw clutch on the driveshaft,
causing the clutch to snap into engagement, as shown in Fig. 13.12. Depending on
which trip lever is raised, the respective clutch stays engaged long enough to complete
one turret index, one stock feed, or one spindle clutch shift.
The trip dogs are bolted in T slots in the sides of the dog carriers and are adjustable
circumferentially. The clutch is engaged when the toe of the trip-lever dog approaches
the high point of the trip dog on the carrier.
The chuck trip dogs normally are set to trip the chuck and feed clutch as the cutting-
off tool clears the stroke on its return stroke. On multiple-geed cuts or on second-
operation jobs other requirements dictate the timing.
The spindle-reversing (or speed-change) dogs are often set to shift the spindle clutch
during a turret index. A more critical setting would be to reverse the spindle to back out
a tap. In this case, the dog is set to trip the clutch when the turret lead lever is on the
high point of the tapping to be on the cam and the turret slide about to start back.
Turret trip dogs are normally set to trip the clutch so that the index is completed just
before the lead-lever roll contacts the beginning of the next rise on the lead cam. On
some jobs, turret indexes are called for during low-point dwells on the lead cams,
particularly when not all the turret stations are in use.

Trip-lever abutments Side pressure from the clutch spring on the eccentric teat of
the screw in the trip lever is taken by the abutments in the driveshaft brackets (see Fig.
13.11). A hardened screw in each trip lever, bearing against the hardened abutment,
provides a means of adjustment.

Second-operation work When the machine is to be used for second-operation work,


it is necessary to arrange the chuck so that the first time the chuck trip lever is raised
the chuck will remain in an open position, allowing time to insert a piece of work, and
the second time the chuck trip lever is raised the collet will close. The chuck clutch
makes two revolutions to complete the feeding of stock, and consequently stopping
this clutch at each revolution will cause the chuck to complete one-half cycle at each
movement of the trip lever. This is accomplished by lowering this shoe a sufficient
amount causes the eccentric screw to engage the chuck clutch at every revolution.

Cross slides Both front and rear cross slides are mounted on hardened and ground
replaceable steel ways, of T-shape cross section, fastened to the machine bed.
Tapered gibs, with adjusting screws and locknuts, provide for side clearance. These
gibs should be set as snug as possible to still allow smooth slide action without any
tendency to stickiness. Gib adjustment should be checked about once a month during
normal operation.
Automatic Screw Machine 841

Two steel straps bolted to the underside of the slides are fitted with a small clearance,
as shown in Fig. 13.14. If this becomes excessive because of wear after a long period,
the original clearance is restored by removing the straps and grinding a sufficient
amount from the upper outer surface (surface A, Fig. 13.14) of the straps to result in
a maximum of 0.025 mm clearance.

Fig. 13.14 Section through cross slides (Brown & Sharpe Mfg. Co.)

Cams installed on the right end of the front camshaft control the feed of the front and
rear cross slides (Fig. 13.8). Holes at the front of the cam levers permit the operator to
insert a pin wrench and lift the levers for manual operation of the cross slides during
setting up.
Both cross slides have a rack C with a lever and segment which imparts movement
in accordance with the outline of its respective cam (see Fig. 13.15). When a slide is
advanced, a coiled spring D is compressed to return the slide quickly and keep the
cam-lever roll bearing on the periphery of the cam. Both slides have an adjusting
screw E furnished with a dial F. A 3.177 by 28.597 mm cold-rolled-steel shearing pin
fastens the dial on the adjusting screw to prevent other parts from being broken in
case of a jam. This arrangement provides means for adjusting the tool posts H and
J toward or away from the work without changing the position of the cam levers and
ensures cutting to exact depth with a forming-tool. An adjustable screw at K acts as
a positive stop; when setting up, the slide is adjusted by the dial adjustment so that
the piece is formed small by approximately 0.1 to 0.12 mm and then the positive stop
screw is set to bring the work to size.

H J

F D Back
E
E

C F
K Front

Fig. 13.15 Section through cross slides and tool posts (Brown & Sharpe Mfg. Co.)

Turret The turret (Fig. 13.16) is indexed from one tool position to the next by a
Geneva movement. Power for this is taken from the driveshaft through a clutch and
gearing.
The turret is provided with six evenly spaced holes into which the shank of the turret
tool is placed and securely clamped with a clamp bushing and clamp screw.
842 Tool Design

Fig. 13.16 Turret and turret slide (Brown & Sharpe Mfg Co.)

The turret is automatically locked at each tool position by the locking pin, which
engages bushings set in the rear face of the turret directly in back of each working
position. Adjustable dogs on turret dog carriers start the turret-indexing mechanism.

Turret indexing The adjustable trip dogs on the turret dog carrier are set to lift
the turret trip lever thereby engaging the indexing mechanism through the turret
clutch (see Fig. 13.17). When setting up with a handwheel, the driving shaft must be
operated through the complete cycle for each tool position. Single or double indexing
is accomplished by simply repositioning the turret locking-pin operating lever (Fig.
13.17), adding or removing a second turret change tool as required.
Single indexing means that the turret indexes to each of the six tool positions; with
double indexing the turret indexes to every other tool position.
The turret is locked in position by a hardened locking pin, which is automatically
withdrawn by the locking-pin cam movement when the indexing mechanism starts to
operate. When the locking-pin operating lever (Fig. 13.16) at the top of the turret slide
is pulled forward toward the front of the machine, it permits the operator to rotate the
turret by hand through one or all six tool positions.

Turret slide The turret slide (Fig. 13.16) upon which the turret is mounted slides on
hardened and ground replaceable steel ways fastened to the machine bed. A tapered
gib with adjusting screws provides adjustment to compensate for wear.
The slide is moved toward or away from the spindle and worked through the lead
cam, lead lever, connecting link, segmented lever, and withdrawal cam, as shown in
Fig. 13.18. The segmented lever operates the adjustable turret-slide rack fastened to
the underside of the slide.
When deep drilling operations are performed, it is necessary to withdraw the drill
one or more times to clear the hole of chips and cool the drill. This withdrawal and
return of the turret slide is done while the index of the turret from one tool position to
the next has been halted by action of the cam (designed for the particular job) on the
rear end of the lead camshaft and a lever which holds the turret driving gear (drive for
turret rotation) out of clutch engagement. With the driving gear disengaged, the turret
withdraws and returns without indexing.
Automatic Screw Machine 843

Initial position of roll and disk Initial position of roll and disk
0.254 to 2.54 mm for six-hole turret 0.254 to 2.54 mm for six-hole turret
0.254 to 3.813 mm for eight-hole turret 0.254 to 3.813 mm for eight-hole turret
Change
roll
Change
roll

Locking-pin
cam Locking pin
cam extension
Locking-pin cam
(a) (b)

Pull-off and Push-on of Turret Tools


size in inches
Six-hole turret Eight-hole turret*
Pull-off Push-on Pull-off Push-on
Position 1, single index 59.58 32.92 58.78 41.30
Position 2, double index 42.90 0.79 40.51 1.59
*When specified

Fig. 13.17 Turret Mechanism for single or double indexing: (a) Single indexing, position 1,
showing position of roll and locking-pin cam (b) Double indexing, position 2,
showing position of rolls, locking-pin cam, and locking-pin-cam extension (Brown
& Sharpe Mfg. Co.)

As the withdrawal cam (Fig. 13.18) rotates into a cam drop, it releases pressure on
the linkage operated by the lead cam, thereby allowing strong spring pressure with in
the turret-slide ways to effect the withdrawal. This withdrawal and return of the turret
takes 0.5 sec.
When setting up to use the rapid pull-out mechanism the turret trip dogs must be set
so that the indexing mechanism (for turret rotation) is actuated only when the pull-out
cam-lever roll (Fig. 13.11) is in a dwell position on the pull-out cam; otherwise only
partial indexing of the turret may result.

Turret-slide rack The adjusting screw (Figs 13.16 and 13.18) in the end of the
turret-slide rack is provided to allow the turret slide to be advanced or withdrawn
within limits of 25 mm without changing the position of the lead-lever roll on the lead
cam.

Constant-stroke turret slide The turret slide has a constant-stroke design. The
withdrawal stroke of the turret slide during turret indexing is always the same,
regardless of its position at the beginning of the index. No stop pin limits the stroke, as
on earlier machines. The position of the turret is always determined by the lead and
withdrawal cams.
The constant stroke is determined by the withdrawal cam, which gives a maximum
movement at the turret slide of 59.58 mm (when single indexing with six-hole
844 Tool Design

turret) before turret rotation starts and a total stroke to the rear of 63.55 mm during
indexing.

Adjusting Segmented lead lever


screw
Rack

Trunnion block
Not to exceed 0.001 and rolls
Clamp
bolts
Adjusting
screw

Lower
Withdrawal screw
Lead lever cam Turret slide
lever plate
Link

Fig. 13.18 Section showing turret-slide drive (Brown & Sharpe Mfg. Co.)

With the rapid-pullout mechanism the


turret-slide withdrawal is a constant
63 mm; as there is no turret rotation, the
full throw of the withdrawal cam can be
utilised.
If greater turret-slide movements than
those mentioned are necessary to
clear the turret tool on an index or to
provide for a rapid pullout, the additional
movement must be provided with a drop
on the lead cam.

Positive stop screw for turret slide A


positive stop screw is provided in the
Fig. 13.19 Top turret slide with cover removed
end of the turret slide for maintaining
(Brown & Sharpe Mfg. Co.)
extreme accuracy of depth when
exceptionally close tolerances are specified. This stop can be utilised at only one
turret position and in conjunction with the highest lobe on the lead cam (The stop can
be used for more than one position if all the lobes using the stop are the same height
as the highest lobe on the cam, within 0.05 mm).
The turret is mounted in two opposed-taper roller bearings. Adjusting nuts, shown
in Fig. 13.19, are provided to compensate for wear after long periods of operation,
making it possible to keep looseness, both radial and axial, to a minimum. Turret
Automatic Screw Machine 845

lift or play should be checked at least


once every six months. If the machine
is operated more than 8 hr per day with
fast cycles, checking should be more
frequent.

Rear upper slide This slide gives an


additional tool position (see Fig. 13.20).
Usually used for a cut-off tool, it leaves
both cross slides free for other tooling.
Cutting off can be accomplished with the
spindle running either direction.
The slide shear-pin lever (Fig. 13.10)
is provided with a shearing pin in case
of overload or jamming. The slide cam
lever has a front hole for inserting a
handle to operate the slide by hand Fig. 13.20 Rear upper slide, showing positive
when setting up. A tapered gib, held by stop and gib adjustment (Brown
adjusting screws, provides means of & Sharpe Mfg. Co.)
keeping the slide and slide way in proper
adjustment.
The link adjusting screw (Fig. 13.10) provides a means of adjusting the tool point
relative to the periphery of the rear-upper-slide cam.
The slide stop screw (Fig. 13.20) is a positive stop against the advance of the slide
beyond the point determined by the cam lobe, thereby assuring uniform advance of
the slide tool.
When the cutting-off blade is reversed to cut in either spindle direction, the tool wedge
B and blade must be reversed and the wedge adjusted for the correct clamping
pressure (see Fig. 13.21).

13.3 TOOLING FOR AUTOMATIC SCREW MACHINES


The tools discussed in the following pages are basically designed for use on Brown
and Sharpe machines, but they can be used with maximum efficiency on lathes,
multispindles, and other machines and serve well as examples of tools for all types of
automatic screw machines.

Plain hollow mill The pain hollow mill is a turning tool held in an adjustable holder
for turning a diameter from the turret (see Fig. 13.22). Its advantage is that it has
three cutting edges and therefore removes a larges volume of material in a short time.
The finished turned diameter passes through the hollow mill, which provides added
support and thus prevents long, small diameters from whipping.
The hollow mill is the only turning tool available that will turn a diameter from the
turret and overlap a narrow form tool at the same time. This is particularly desirable
in making screws, where the length of body is out of forming ratio and the place has
a neck to allow the thread to be cut close to the head. To produce the best quality
of screw, the neck should always be formed before the thread is cut. As the necking
846 Tool Design

0° 30°
D

Fig. 13.21 Diagram for rear-upper slide adjustment (Brown & Sharpe Mfg. Co.)

operation can be overlapped in a plain hollow mill, a screw made in this manner will
be produced much faster than a job laid with any other type of turning tool.
The plain hollow mill has two main disadvantages: (i) Since each mill will turn only one
size, a company using hollow mills exclusively for turning would soon build up a rather
large inventory, (ii) A cutter-sharpening fixture must be used to sharpen the three
cutting edges in the same plane to ensure cutting efficiency. Many screw-machine
shops lack this equipment.
The adjustable hollow mill, shown in Fig. 13.23, is primarily used on hand machines.
The type is sometimes listed as being adaptable to the automatic screw machine but
generally it has not found favour in this application.

Box tools The box tool shown in Fig. 13.24, is available with carbide backrests
and is recommended for taking light facing cuts. The depth of cut for this type of tool
should not exceed 0.50 mm except for free-machining brass, where medium to heavy
cuts can successfully be taken. Generally, this tool is considered superior to the roller-
backrest box tool on this material.
The roller-backrest box tool, shown in Fig. 13.25, is used for medium or heavy cuts
where a single finishing cut is desired. To accomplish the purpose of the rollers,
the blade should take a chip heavier than 0.5 mm in depth but not exceeding the
Automatic Screw Machine 847

Fig. 13.22 Plain hollow mills (Brown & Fig. 13.23 Adjustable hollow mill (Brown
Sharpe Mfg. Co.) & Sharpe Mfg. Co.)

(a) (b)

Fig. 13.24 Box tools with carbide V rests. Right-hand tools with (a) two blades and (b) one
blade. (Brown & Sharpe Mfg. Co.)

diameter of the stock divided by 5. This tool is satisfactory on all materials except
free-machining brass, for which the box tool is rated superior because the pressure of
the cut is so great against the rollers that it results in extrusion and elongation of the
turned diameter may occur.

Balance turning tool Often seriously underrated, this tool is a one-piece drop
forging with a pair of heavy blades securely clamped and is capable of removing more
material in a given time than any other tool at the operator’s disposal (see Fig. 13.26).
The balance turning tool can be set in two different ways.

Fig. 13.25 Roller-backrest box tool Fig. 13.26 Balance turning tool (Brown
(Brown & Sharpe Mfg. Co.) & Sharpe Mfg Co.)
848 Tool Design

Roughing This is accomplished by setting one blade to cut to size, bringing the
second blade in contact with the turned diameter (also to size), and then advancing
the tool with the turret handle so that material fuzz is found on both blades. When
this condition exists, both blades are cutting to size and removing equal amounts of
material. A balance turning tool should smoke during its operation, a condition that
does not apply for most other screw-machine tools. If the tool does not smoke, its full
efficiency is not being realised.
Combination roughing and finishing tool One blade is set for roughing and should
remove about 80 per cent of the total material. The second blade is set slightly behind
the roughing blade, a distance equal to at least the feed per revolution, and is set to
the finish size. When large volume of material are to be removed, the balance turning
tool should always be considered.
If a smooth, accurate finish is desired on a part, the piece may be roughed with a
balance turning tool and then followed by a finishing cut with a box tool. This excellent
combination results in very good finish and long tool life between sharpening of blades.
For tolerances on the turned diameters of 0.025 mm or less, this combination is highly
recommended. The balance turning tool is commonly used for turning thread-blank
diameters. Since the balance turning tool is normally considered a roughing tool, the
die head becomes the finishing tool.

Floating and adjustable holder The adjustable


holder, shown in Fig. 13.27, is popular spotting,
centreing, and holding drills, countersinks, or
counterbores. The floating or adjustable principle,
available only during the set-up period, corrects
misalignment between the spindle and the turret
hole, assuring concentric internal operation.
When maintaining close ID tolerances, the hole usually
drilled 0.152 to 0.254 mm undersize and then either
redrilled or reamed to hold finish diameter. When
Fig. 13.27 Adjustable holder
machining abrasive material to tolerances closer than
(Brown & Sharpe
0.025 mm the job when first set-up with a high-speed-
Mfg. Co.)
steel stub screw-machine reamer will hold size for
only a short time and then wear out of tolerance.
To overcome this problem, carbide-tipped reamers in the larger sizes and solid-
carbide reamers in small sizes are suggested. Because carbide is as brittle as it
is hard, a high breakage often results when the reamer is locked up solid with an
adjustable holder. This problem can be solved by using a floating reamer holder,
which is similar in appearance to an adjustable holder. It has a floating body held in
place with a neoprene O ring, allow sufficient float parallel to the spindle axis.
It is common practise when producing holes on the automatic screw machine to ream
to close tolerance. If concentricity is a factor, the hole must be bored; it is standard
practise to bore a hole for concentricity only and not for finish. If a hole must be
concentric and also accurately sized, standard practise would be to spot or centre-drill,
bore, and then team. The reamer follows the hole and does not alter the concentricity
produced by the boring tool. The boring tool resembles an adjustable holder but has a
Automatic Screw Machine 849

radial movement which can be governed with


a vernier adjusting screw to gibe the operator
accurate control of final size adjustment.

Swing tool For taper turning, straight


turning behind a shoulder, or contour turning
the swing tool is an old standby (see Fig.
13.28). Because the toolholding arm swings
from a fulcrum screw, a slight error results,
making it impossible to have the taper Fig. 13.28 Turning swing tool (Brown
perfectly straight in line. & Sharpe Mfg. Co.)

Slide tool This tool is designed to correct the errors of the swing tool. It not only
produces a straight taper, but since the radial travel of the adjustable guide must be
set approximately twice the travel of the cutting edge of the swing tool, it can produce
a maximum taper with nearly twice the angle of which the swing tool is capable. The
slide tool can be used for radial turning the face of a part, which is not practical with a
swing tool. Figure 13.29 shows a slide tool, and Fig. 13.30 shows its use.

Fig. 13.29 Slide tool (Brown & Fig. 13.30 Using the slide tool to turn tapers
Sharpe Mfg. Co.) (Brown & Sharpe Mfg. Co.)

Die holders Most of the external threads on screw machines are cut with self-
opening die holders, an excellent tool for medium and coarse threads. The button
die holder is in moderate demand, particularly where odd threads are concerned, but
because of the difficulty of obtaining good dies and the problems of sharpening, the
self-opening die holder is gaining in favour.
For fine, accurate threads on the automatic screw machine, a solid adjustable die
with insert chasers is preferred. The powerful retaining spring in the self-opening die,
although no problem on medium or coarse threads, becomes a serious problem in
starting fine pitches on the thread blank and often produces a tapered-pitch-diameter
screw.

Tap holders and taps Die holders and tap holders are furnished in nonreleasing and
releasing models. The nonreleasing type is commonly used on the automatic screw
machine when reversing the spindle is a mechanical operation.
850 Tool Design

The releasing holder is generally used on the hand screw machine where reversing
is a manual operation, since releasing holders allow for human error. In rare cases,
the releasing die or tap holder must be used on the automatic screw machine when
the time cycle is very slow or the pitch of the thread is very coarse because either
condition or a combination of both results in such a steep rise and drop it is impossible
to get a practical thread lobe on the cam.
When rough and finish taps are used on a screw-machine job, the first, or roughing
tap is held in a regular nonreleasing tap holder. The second, or finishing, tap should
always be held in a finishing tap holder. The main difference between a finishing tap
holder and the standard tap holder is the addition of one or two pin plungers held
under spring tension to give a buffer action and allow the tap to float horizontally.
The finishing holders allows the finishing tap to synchronise with the roughing thread
previously cut.
A satisfactory method of cutting small, coarse threads is to remove two-thirds of the
material with the roughing tap and one-third of the material with the finishing tap. This
is controlled by the pitch diameter on the roughing tap, and the formula is standardised
with tap manufacturers. It is therefore only necessary to order rough and finish plug
taps from the manufacturer.
All taps used on the automatic screw machine should have ground threads. Cut
thread taps have varying degrees of distortion due to heat treatment; this increases
the torque during the cutting operation and may encourage tap breakage. The ground
thread taps supplied by the manufacturer are made in a range sufficient to cover the
popular tap classes.

Angular cutting off The angular cutting


off shown in Fig. 13.31, is a satisfactory
way of severing a piece from the bar when
a blunt point is desired, such as the ends
found on a drill or tap blank. This tool is held
in the turret and is indexed into position
on the machined part. For best results,
a bushing support in the tool is used to
reduce whipping.
A dwell on the lead cam holds the tool Fig. 13.31 Angular cutting-off tool
steady in a longitudinal position, and a (Brown & Sharpe Mfg. Co.)
fixed or adjustable guide on the front slide
in contact with a screw at the bottom of the tool transfers a radial thrust to an angular
movement. The cutting tool following this course transfers all the cutting torque to the
collet, and the piece to be made just “goes along for the ride”. The part is severed with
a smooth cone and a cut-off scar not to exceed 0.076 to 0.152 mm diameter. This is
acceptable in standard screw-machine practise.
This tool has an adjustment with the centreline ranging from 25 to 45°, producing
a cone of from 50 to 90° included, which satisfies the majority of parts of this type.
Should other angles be required, the tool can be specially built to cover other ranges.
Another popular application of this tool is to cut a grinding neck in the corner of a piece
where a face grind and cylindrical grind are required.
Automatic Screw Machine 851

Circular tools The tool posts supplied as standard


equipment on the automatic screw machine are
designed for circular tools (see Figs 13.32 and
13.33). Circular tools are popular because they are
economical to produce and have long cutting life.
For radial or longitudinal surfaces on screw-machine
parts, the circular tool does a satisfactory job, but if
a series of precision angles is required on a part, the
circular tool, when sharpened parallel to the radius
but offset for clearance, introduces an error of radial
difference and the resulting angle is not flat but has
a slight hollow. Fig. 13.32 Tool post for circu-
lar tools (Brown &
Since this cannot be corrected in the circular tool, Sharpe Mfg. Co.)
where flatness is essential, universal tool posts for
square tools must be used, as shown in Fig. 13.34. The universal tool post is widely
used for cutting tools with carbide inserts. When a circular tool is designed with carbide
inserts, the maximum number of inserts should be held to three to allow a surface on
the perimeter of the tool for secure clamping by the hook bolt. The clearance angle,
on the cutting edge of a circular tool is built into the circular tool holder and ranges
from 8° 13’ to 9° 36’, depending upon the size of the machine.

Fig. 13.33 Tool post for circular tool with tool in position (Brown & Sharpe Mfg. Co.)

Fig. 13.34 Tool posts for square tools (Brown & Sharpe Mfg. Co.)
852 Tool Design

The clearance angle which is perfectly satisfactory for high-speed-steel and cast-alloy
tools is too great for carbide, for which the recommended clearance angle is from 4 to
6°. By making two or more duplicate carbide-tipped flat tools with the proper clearance
angle, the carbide-tool application is more satisfactory. One tool can be used until it
requires sharpening and then be replaced by another tool, allowing the machine to
continue production while the first tool is being resharpened.
On the screw machine, most turning blades are tangential, with the cutting edge on
the end, thus allowing the blade to be rocked back and forth or mover in and out
tangent to the cut. This arrangement gives fine adjustment and makes it easier for the
operator to correct for errors in turned diameters.

Knee tool The knee tool is an exception to the statement above and cuts radially in a
manner similar to that of a lathe tool, resulting in excellent chip clearance and narrow
tool (see Fig. 13.35). The latter feature reduces the clearance time between the turret
and the cross-slide tools.
The knee tool has no support and is limited to the removal of scale from hot-rolled
material, to short, light cuts close to the chick, or to facing a piece for controlled length
and finish. It should not be selected for long-diameter turning or when high finish and
close tole-rance must be held.

Centreing-and-facing tool This is a combination tool. When feeding stock to a turret


stop up to the recommended maximum length, a feeding-stock tolerance of ± 0.076
mm can be held. If the requirements of the workpiece call for a tolerance smaller than
this, the part is usually fed about 0.254 mm longer and the extra material faced off
from the turret. A tolerance of ± 0.025 mm can be held in this manner. If the piece is
short and can be faced with form tool, a tolerance of ± 0.012 mm be maintained.
The centreing-and-facing tool is equipped with a centreing drill as shown in Fig.
13.36. If the workpiece requires a drilling operation, this combination tool will centre
or spot the piece and face it to the required length. The centreing-and-facing tool has
sufficient movement of the blade to allow facing solid stock.

Pointing tools Pointing tools are of two different designs. The circular pointing
tool, shown in Fig. 13.37, has a circular-formed tool similar to cross-slide tools. An

Fig. 13.35 Combination right- and left- Fig. 13.36 Combination right- and left-hand
hand knee tool (Brown & adjustable centreing-and-facing tool
Sharpe Mfg. Co.) (Brown & Sharpe Mfg. Co.)
Automatic Screw Machine 853

economical tool with a very long life, it can be used right- or left-handed and requires
no clearance from the cross slide, therefore, its operation can be overlapped with
the cross slide. Since this tool has no support, its use is limited to short piece. As a
general rule, the circular forming tool should not be used if the part projects from the
turret more than four times the stock diameter.
For long pieces, where the stock projects more than four diameters, the pointing
tool shown in Fig. 13.38 should be used. This tool has a V support just ahead of the
blade, holding the workpiece with sufficient rigidity to prevent chatter and provide
good tolerance. The V backrests furnished with the tool are chrome carbide. The
blade furnished with the tool can be ground to suit the required contour on the end of
the piece. To furnish sufficient coolant to the cutting edge, the blade must be used in
a horizontal position, and this necessitates clearance from the front slide.

Fig. 13.37 Pointing toolhoder for circu- Fig. 13.38 Pointing tool with carbide V
lar tools (Brown & Sharpe backrest (Brown & Sharpe
Mfg. Co.) Mfg. Co.)

Knurling tools Several types of knurling


tools are available for automatic screw
machines. Figure 13.39 shows the turret
adjustable knurl holder, which will generally
produce the best quality knurl. The statement
usually associated with this holder reads,
“The knurls are mounted in swiveling holders
adjustable to any angle to produce straight,
spiral or diamond knurling using ordinary
straight rolls.”
Fig. 13.39 Turret adjustable knurl
Up to a point, this statement is true. A straight holder (Brown & Sharpe
knurling roll set at an angle will produce Mfg. Co.)
a good spiral or male diamond knurl. The
straight knurl roll in contract with the knurl blank at an angle will extrude the material,
and since the axis of the knurl and the stock are not parallel, a beneficial burnishing
effect results. The bottom of the knurl extends in a radial plane, as shown in Fig.
13.40. Therefore, if the work should spring slightly away from the centreline, a full
knurl will still result. If a straight knurl on the piece is produced with a straight knurling
roll, the knurl advances on the blank with an effect similar to that of a snow plow.
The cuts of the knurl are gouged out of the material, producing a rough knurl with a
minimum of build-up.
854 Tool Design

Fig. 13.40 Position of knurl rolls on Fig. 13.41 Knurl rolls set straight pro-
turret adjustable knurl duce chewing effect with any
holder (Brown & Sharpe movement of the workpiece in
Mfg. Co.) the horizontal plane (Brown
& Sharpe Mfg. Co.)

If the piece is long, any movement in a horizontal plane of the knurl blank will bring the
knurl blank out of contact with the knurling roll, resulting in a serious chewing effect,
as shown in Fig. 13.41.
To produce a straight knurl, it is better to use a spiral knurling roll in the holder and set
the roll at the angle of spiral in relation to the part (see Fig. 13.42). With this set-up,
all the benefits of a straight knurling roll producing a spiral knurl are retained. A spiral
knurl ranging from 20 to 30o is the most satisfactory. Two spiral knurls of the same
rotation should be used, not one right and left hand.

Fig. 13.42 Position of spiral knurling roll (Brown & Sharpe Mfg. Co.)

In spite of the excellent work done by rolling mills in rolling round stock, the stock is
not perfectly round. It is slightly elliptical or egg-shaped. When turret knurling on stock
size, this error in the periphery of the stock is exaggerated in the finished knurl.
If a near-perfect knurl is required, the knurl blank should always be turned first to give
roundness and concentricity with the machine spindle. On mild steel, an excellent
spiral or diamond knurl can be produced with a feed of 0.203 mm/rev on and 0.305 mm/
rev off. This is somewhat lower than the standard knurling feed table but is capable of
producing a diamond knurl that will stand inspection under a 10x magnifying glass.
A second type of knurl holder, the side knurl holder, (Fig. 13.43) can be held in either
the front or back cross slide but is usually held in the latter. Generally employed for
knurling behind a shoulder and close to a chuck, this holder is versatile enough to
produce a straight, spiral, or diamond knurl on the workpiece, depending upon the
knurling roll employed. A male role will produce a female knurl on the workpiece,
and vice versa. The side knurl holder is easy to set-up, and the cam is easy to make.
Unfortunately, when the side knurl holder is used, it eliminates a circular tool.
Automatic Screw Machine 855

Fig. 13.43 Side knurl holder for cross slides (Brown & Sharpe Mfg. Co.)

The top and bottom knurl holders for cross slides,


shown in Fig. 13.44, are mounted on a hub
integral with the cutoff tool, producing a sort of
have-your-cake-and-eat-it-too condition. The top
and bottom knurl holders pass over and under
the stock, respectively, with a truncated lobe on
the back slide. The back slide then advances into
the cut-off position and severs and severs the
piece in the usual manner.
This produces a very satisfactory knurl but is
practical only when the knurl on the piece is close Fig. 13.44 Top knurl holder for
to the chuck, as the supporting hub for the tool cross slides (Brown &
controls the relation of the knurl to the end of the Sharpe Mfg. Co.)
piece; if the requirements bring the knurl well away from the chuck, a very wide cut-
off tool results. The hub involves additional work on he cut-off tool, which adds to the
expense.
At the completion of the cut-off operation, the back slide not only must drop back to
clear the stock with the cut-off tool but must clear the top of bottom knurl holder as
well. This clearance is time consuming and is approximately five-hundredths of cam
surface in most cases. On the older screw machines that have no vertical slide and
no provision for mounting one, the top and bottom knurl holder has been a realiable
standby. Later model machines have vertical
slides in most cases, and the current line of
automatic machines has vertical slides as
standard equipment, due to popular demand.
Because of this trend, the top and bottom
knurl holders are losing popularity.
The last knurling tool to be described is the
knurling swing tool, shown in Fig. 13.45,
the most versatile of the knurling tools. It
can be used for knurling behind shoulders
and for knurling a section at some distance
from the end of the workpiece. It is also used Fig. 13.45 Knurling swing tool (Brown
for longitudinal knurling and for light thread & Sharpe Mfg. Co.)
856 Tool Design

rolling. The knurl is held in the swinging arm, which is pivoted on the main body of the
tool. The arm is activated by a fixed or an adjustable guide mounted on the front cross
slide. The shank is designed to hold the backrest on a special support which car be
clamped into position by a setscrew.

Summary

Automatic screw machine was developed from lathe for economical manufacture of
small screws and bolts. Other important parts that can be produced in these machines
are shafts, compressor nuts, threaded retainer rings, knobs, tapered plugs, spark-plug
shells, and artillery shells. This chapter deals with different aspect of this specialised
machine.
• The principal features of the automatic screw machine are automatic stock
feeding and operation of all cutting tools by cam action.
• Automatic screw machines may be single spindle or multiple spindle type.
• An automatic screw machine uses spring collet for automatic bar feeding.
• The turret of an automatic crew machines is indexed from one tool position to
the next by a Geneva movement.
• The adjustable trip dogs on the turret dog carrier are set to lift the turret lever
thereby engaging the indexing mechanism through the turret clutch.
• The turret slide upon which the turret is mounted slides on hardened and
ground replaceable steel ways fastened to the machine bed.
• Automatic crew machines require specially designed tools for better perfor-
mance. Some examples of such tools are plain hollow mill, box tools, slide
tools, die holders, tap holder and taps, knee tool, etc.

Questions
1. What is the advantage of using a swing stop for stock?
2. How is the drive from the spindle motor transmitted to the spindle?
3. What is the purpose of the driveshaft?
4. What is the operating speed of the driveshaft?
5. How are the driveshaft and chuck and feed cam powered?
6. What is the speed range and how is the speed of chuck and feed cam drive
changed?
7. What determines the selection of the four feed change gears used on the
Brown and Sharpe Ultramatic Screw Machine?
8. What variation in stock size is allowable for satisfactory chucking?
9. What is the purpose of the feed-tube bushing?
10. What is the purpose of the chuck camshaft?
11. What time is required to feed the stock?
12. What protection is built into the feed slide?
13. What controls turret indexing?
14. At what stage in the machining operation should the chuck trip dogs trip the
chuck and feed clutch?
Automatic Screw Machine 857

15. What type of movement is used to index the turret?


16. What is the purpose of the rear upper slide?
17. What tool is usually used in the rear upper slide?
18. What are the major advantages of the plain hollow mill?
19. When using the roller-backrest box tool, what minimum depth of cut will cause
the roller to function properly?
20. What is the major advantage of the balance turning tool?
21. How can the operator tell when the balance turning tool is cutting to full
efficiency?
22. What are the two different methods of setting the balance turning tool?
23. How is the balance turning tool used when a smooth, accurate finish is desired
on the workpiece?
24. What is the purpose of the adjustable holder?
25. What is the purpose of the floating reamer holder?
26. What machining procedure should be followed when a hole requires close
tolerance, good finish, and concentricity?
27. Why is it impossible to have a perfectly straight taper when using a swing
tool?
28. What tool is used to correct the errors of the swing tool?
29. Why is a solid adjustable die preferred when cutting fine, accurate threads on
the automatic screw machine?
30. What is the difference between a finishing tap holder and a standard tap
holder?
31. Why should all taps used on screw machines have ground threads?
32. What is the angular range of the cone produced by the angular cutting-off
tool?
33. How is the clearance angle provided on the circular tool?
34. Why is the knee tool generally limited to light cuts?
35. Why should a pointing tool with V backrests be used when the stock projects
more than four diameters from the collet?
36. What is the best method of producing a straight knurl with the turret adjustable
knurl holder?
37. Why should the knurl blank be turned before knurling when a near-perfect knurl
is required?
38. Where is the side knurl held in the automatic screw machine?
39. What is the disadvantage of using the side knurl holder?
40. How is the knurling arm of the knurling swing tool activated?
858 Tool Design

References
Books
• Brown & Sharpe Mfg. Co.: “Operation and Maintenance of the No. 00 Ultramatic Screw
Machines,” North Kingstown, RI.
“Operation and Maintenance of the Nos. 2 and 3 Ultramatic Screw Machine,” North
Kingstown, RI.
“Cam and Tool Design,” North Kingstown, RI.
“Brown and Sharpe Cam and Tool Design Tables,” North Kingstown RI.
“The Shop Tool Manual,” North Kingstown, RI.
• National Acme Company, “Acme-Gridley Multiple Spindle Bar Machine Manual,” Cleveland,
Ohio, 1961.
Catalogues
• Acme Industrial Co., Chicago, III.
• Brown & Sharpe Mfg. Co., North Kingstown, RI.
• Cone Automatic Machine Co., Windsor, Vt.
• Hardinge Brothers, Inc., Elmira, NY.
• Carl Hirschmann Co., Inc., Manhasset, NY.
• New Britain machine Co., New Britain Machine Division, New Britain, Bonn.
• Traub U.S.A., Inc., Plainview, NY.
Periodicals
• Automatic Machjning, Rochester, NY.
INDEX

A Angularity 391
Abbreviations commonly used in tool Animal oils 263
drawings 9 Annealing 157
Abbreviations used on tool drawings 14 Antifriction bearings 156
Abrasion 234 Approximate Steel-Hardness Conversion
Abrasive-cloth disks 54 Numbers 142
Abrasive-clothe rotary tools 52, 54 Arbor equipped with expanding shoe 570
Abrasive sticks 43 Area relief 451, 506
Accuset 777, 778 Arkansas stones 44
Acid etching 117 Assembly dies 689
Adjustable boring toolholder 301 Assembly fixtures 537
Adjustable chip breaker 293 Automatic button die stop 618
Adjustable hand reamer 356 Automatic changer 815
Adjustable holder 848 Automatic die stops 660, 662, 661
Adjustable hollow mill 847 Automatic feeding 599
Adjustable locators 443 Automatic fixtures 546
Adjustable punch and die set 667 Automatic gauges 423
Adjustable reamers 357 Automatic grinding machines 424, 769
Adjustable rest pins 441 Automatic lathes 769
Adjustable trip screw 660 Automatic loading fixtures 557, 558
Adjustable vise 779 Automatic operating mechanism 836
Air bend dies 681, 682, 683 Automatic screw
Air bending 681, 687 machines 769, 830, 848, 849
Aircraft extension drill 339 Automatic screw machining 267
Air cushions 691, 692 Automatic stops 656, 660
Air dies 687 Automatic tool changers 809, 814
Air-hardening cold-work steels 151 Average mechanical properties of various
Air-slide feeds 602 alloys 716
Air vent 699, 700 Average specific energy 201
Allotropy 162 Axial rake 185, 186, 330
Aluminium oxide abrasive 44 Axial rake angle 320
American National Form (ANF) 396 Axial relief 322
Angle 187, 353 Axial relief angle 320
Angle cutters 316 Axial run out 318
Angle of inclination 280, 362 Axis 336
Angle of thread 361 Axis of tap 361
Angle-type fixture 567 B
Angular 290 Babbitt 126
Angular clearance 608 Backlash 80, 312
Angular cutting-off tool 850 Backlash eliminator 312
Angular flutes 318, 325 Back-pin die sets 614
860 Index

Back-pressure air gauge 413, 414 Bolster 617


Back rake 187, 280 Bolster Plate 692, 595
Back-rake angle 273, 278 Boring 272
Back rate 274 Boring and turning 267
Back taper 323 Boring-bar 87, 552 555, 555
Back-up plates 631 Boring-bar cutters 300
Balance turning tool 847 Boring fixtures 552, 555
Band filling machine 37 Boring head 299
Bar stiffness 255 Boring mills 770
Base of thread 361 Boring operations 294, 295, 552
Basic chip types 190 Boring tool 87, 294, 306, 307, 848
Basic principles of locating 435 Boring-tool geometry 305
Basic principles of location 444 Bottoming 682, 687
Basic tool geometry of preformed Bottoming dies 682, 683
boring tool 296 Bottoming out 683
Bearing 153 Bottoming tap 363
Bearing surfaces 543 Boundary-lubrication 263, 266
Bench block 23 Box and tumble jigs 495
Bend allowance 683, 685, 333 Box jig 495
Bend angle 685 Box tools 846, 847
Bend axis 685 Box-type tumble jig 496
Bending 137, 619 Brass drift 63
Bending dies 680 Brazed-tip tools 283
Bending methods 681 Brazed tools 284
Bending pressures 687 Bridge-type clamping mechanism 287
Bend line 685 Brine 172
Bend radius 685 Broach 559
Best wire size for different thread forms 420 Broaching 267
Bevel-head punch 632 Broaching Fixtures 558, 559
Bismuth alloys 125, 156 Brown and Sharpe automatic screw
Blade 352 machine 830
Blank 252, 253, 694, 700 Brushing 708
Blank diameter 709 BUE chip 190
Blanked 593, 607 Built-in chip-breaker 345
Blank ejection 625 Built-in lubrication system 233
Blank-holder 612, 701, 702, 703, 705 Built-up construction 511
Blanking 591, 624 Built-up edge (BUE) 189, 190, 230, 233
Blanking and Piercing Die Construction 625 Bulging 685, 697
Blanking die 620, 755 Bulging dies 689, 698, 699, 739
Blanking or piercing 604 Bulging nipple into tubing 756
Blanking punches 618, 637 Bulging punch 756
Blanking punch piercing die 620 Bulging shell in split die 756
Blank Size 715 Burnishing 715
Block boring tools 304 Burnishing rollers 156
Bluging punch 756 Burnishing tools 153
Body 318, 691 Burr clearance 91, 92
Body dia clearance 335 Burrs 501
Index 861

Bushing 67, 504 Channel bend 685


Bushing for embedment 510 Channel bending 687
Bushing plate 504 Channel jigs 495, 496
Bushings as guides for reamers 510 Channel strippers 639
Buttons 444, 503 Channel-types strippe 638
Channel-type strippers 640
C
Charge factor 538
Cam and screw clamp assembly 459
Charging method 59
Cam-clamp assembly 458
Chatter 252, 360
Cam clamps 456
Chatter and vibrations 253
Cam relief 322
Chatter marks 252
Cams 836
Checking fixtures 733, 735
Cam shafts 836, 837
Chemical or extreme-pressure lubrication 264
Cam-type latch 462
Chips 56, 184, 236, 537, 542
Cap-screw rest button 507
Chip breakers 286, 287, 289, 290, 603
Cap screws 626
Chip clearance 91
Carbide ball end mill 123
Chip coiling 501
Carbide grade chart 369
Chip control 344, 450, 451, 503
Carbide-tipped machine reamer 356
Chip control in drilling operations 344
Carbide-tipped reamers 357
Chip control with single-point tools 288
Carbide tools 367
Chip disposal 324
Carbon muffle 170
Chip flows 278, 374
Carbon steel 161
Chip formation 184, 188, 329
Carbon-tungsten steels 153
Chip formation in drilling 500
Carburised 153
Chip or burr groove 451
Cast frames 594
Chipping 105
Castings 157, 830
Chip pocket 331
Cast iron 153
Chip removal 294
Cast nonferrous alloys 156
Chip space 323
C clamp 30
Chip thickness 217
Cellulose 154
Chip thickness ratio 192, 559
Cemented carbide cutting tools 368
Chip-tool interface 265
Cementite 161
Chip with a built-up edge 191
Center distance 415
Chisel-edge 333, 335, 341, 415
Center fixtures 546
Chisel-edge angle 336
Centering drill 268, 337
Chisel-edge angle deg 343
Centre drilling or spotting 794
Chromium hot-work tool steels 152
Centre drills 83, 337, 793, 848
Chuck-clutch trip lever 835
Centring-and-facing tool 852
Chuckers 830
Cerromatrix 632
Chuck fork sleeve 835
C-frame fixture 421
Chucking reamers 355, 356
C-frame gauging fixture 422
Chuck parts 153
Chamfer 318, 329
Chucks 561
Chamfer angle 352
Circular tools 851
Chamfering tool 305
Clamping 434, 569
Chamfer length 353, 340
Clamping devices 502
Chamfer relief 353
Clamping forces 452
Chamfer-relief angle 353
862 Index

Clamping hole 612 Continuous chips 189, 323


Clamping methods 502 Continuous chip with 189
Clamp nesting-block 778 Continuous-path milling machine 771
Clamp-on boring cartridge 305 Contour band sawing 41
Clamps 452, 502 Contouring system 770
Clamps should 543 Contour sawing 98
Classification and composition of principal Contour sawing of punches and dies 95
types of tool steels 146 Conventional milling 311, 543
Clearance 318, 687, 689, 697 Conventional or up milling 310
Clearance angles 185, 276 Conventional (up) and 311
Clearance diameter 335, 696 Convex and concave shear 611
Climb and conventional milling 312 Convex cutter (form relief) 318
Climb (down) milling 311 Coolants 263, 268
Climb milling 311, 312, 543 Cooling rate 172
Climb or down milling 310 Coordinate locating system 69
Clip clearance 331 Core diameter 353
Clip load 324 Core drills 338, 339
Cobalt binder 155 Corner angle 187, 329
Coil cradle 600 Corner relief 505, 506
Coined 593 Corner setting 683
Coining 614, 696 Cosine error 25
Coining dies 689, 697 Counterbored 626
Coining (squeezing) 683 Counterbores 632
Collet 84 Counterboring 805
Collet-type drill holders 794 Countersink 57
Collet-type holder 795 Countersinking 56, 57, 695
Combination dies 620, 621 Countersinks 817
Combination feeding and straightening Counterweighted 561
machine 602 Cracking 105
Combination gauges 61, 391 Cradle 600
Combination straightening machine and Cradle fixtures 546, 548
coil cradle 601
Crater 233
Comparative gauging 407
Cratering 233, 235
Comparator 418, 824
Crater-resistant grades 374
Compound die 619
Crater wear 236
Compression 137
Crest 361
Compressive strength 136
Crimping 699, 700
Compressive stress 194
Crimps 700
Concave and convex 316
Critical cooling rate 158
Concave cutter (form relief) 318
Critical range 162
Concentricity 391
Critical temperature 158
Concentric location 439
Cross slides 839, 840
Cone angle 320
Cum-lock long-travel clamp 460
Cone locator pins 442
Cup diameter 709, 713
Conical location 439
Curing 730
Constant-stroke turret slide 843
Curl 694
Continuous 190, 503
Curling 694
Continuous-band filling machine 37
Index 863

Curling dies 689, 695 Die block 56, 63, 98, 625, 626, 689
Cut-and-try method 87 Die-block design 626
Cut off 593, 620 Die-block thickness 627
Cut-off die 617 Die cavity 691, 344
Cutter sweep 353 Die clearance 605, 607, 715
Cutting action in punch and die Die construction 617
operations 603 Die cushion 693
Cutting edges 273, 336 Die-cutting 680
Cutting face 353 Die-cutting operations 591
Cutting fluids 262-264, 267, 271 Die designs 625, 759
Cutting forces 609, 610 Die-draw radii 705
Cutting knives 153 Die files 45
Cutting lips 333, 345 Die-handling equipment 35
Cutting off and parting 592 Die holders 667, 849
Cutting/perforating force 643 Die land 608
Cutting ratio 192 Dielectric fluid 104, 117
cutting speed 250 Dielectric liquid 104
Cutting speeds and feeds 242 Dielectric oil 117
Cutting speeds for milling 244 Die-locating plugs 667
Cutting tools 182, 183, 272 Diemaker’s squares 28
Cuttting Speeds and feeds 240 Die-model duplication 735
Cylindrical cups 710 Die opening 603
Cylindrical land 319 Die-opening factor 687, 688
D Die radius 702
Decarburisation 160, 169, 170 Die rolls 694
Deep box jigs 86 Dies 102, 346, 617, 689, 697
Deep-draw dies 52 Die sections 629
Deep drawing 706, 710, 712 Die sets 601, 612, 613, 618
Deep draws 717 Die-set shoes 153
Deep drilling 842 Die shoe 63, 503, 506, 612
Deep-hole boring 87 Diesinker’s files 45
Demountable boring head 300 Diesinker’s riffler files 46
Densified wood 154 Die space 598
Depth of cut 217, 250 Die springs 643
Depth of draw 709 Die stop 659
Depth of thread 361 Die-stop activating pin 618
Depth-to-width ratio 113 Die thickness 627
Design considerations for urethane dies 750 Differential screw 398
Design procedure 3 Direct or indirect pilots 636
Desirability factors 734 Discontinuous 190
Detail ballons 13 Discontinuous chip 189, 191
Dial-gauge mechanism 409 Disposable insert-type cutting tool 284
Dial indicator 408 Distance 391
Diameter of blank 713 Dog carriers 836, 839
Diamond 155 Dog-point clamps 454
Diamond locating pins 446, 447 Dog-shaft carrier 839
Die 669, 686, 691, 700 Double-action 718
864 Index

Double-action press 597 Drill jigs by casting 750


Double-angle 316 Drill laminating 749
Double-angle milling cutter (profile relief) 317 Drill lips 500
Double-margin drill 796 Drill-point angle 340
Double-or multiple-piloted bar 552 Drill-point geometry 333, 340, 342
Double- or triple-action presses 732 Drill points 341, 345
Double-piloted bar 554 Drill-point wear 346
Double-piloted boring bar 555 Drill presses 552, 799
Double rack-and-point roll feed 601 Drill spilt 345
Double shear on dies 611 Drill walking 792
Double shear on punches 611 Drill web 500
Double tempering 159 Drive 832
Dovetail gauges 398, 399 Drop-hammer dies 732, 739
Dowel 55, 57, 511, 627, 630 Drop hammers 697
Dowel holes 57 Drop-through dies 617
Dowel pin 506, 656 Drum fixtures 546
Dowel-pin reamers 57 Dual-cut drill 340
Drafting and design techniques in tooling Duplicating mills 733
drawings 5 Duplicating models 733
Drafting practise 5 Dynamometer 198
Draw 694
E
Drawability 709
Eccentric-cam contour 456
Draw and form dies 732
Eccentric pin-type clamping mechanism 288
Draw clearance 714
Eccentric relief 105, 323
Draw dies 701, 703, 705, 716, 758
ECEA 277
Drawing 598, 699
Edge finder and indicator 81
Drawing dies 614
Edge finders 31
Drawing force 716
Edge radius 603
Drawing layout 6
Edge radius depth 605
Drawing operations 699
EDM 103, 320
Drawing speed 714
Effective rake 319, 340
Draw radius on die 704
Effective-rake angle 341
Draw ring 718
Elasticity and stiffness 136
Drill 84, 243, 333
Elastic limit 136, 187
Drill bushings 90, 503, 507, 510, 733
Elastoplastics 731
Drill design 344
Electric and electronic indicators 414
Drill diameter 335
Electric furnaces 168
Drilled 632
Electric magnifier 416
Drill fixtures 735
Electro discharge machineg 103
Drill flutes 500
Electronic indicator 416
Drill geometry 500
Electroplating 117
Drilling 56, 333
Eliminator 312
Drilling and reaming 267
Elongation 137
Drilling machines 334
Emboss 696, 717
Drilling operations 333, 345
Embossing 620, 695
Drilling problems 345
Embossing dies 324, 326, 689, 718
Drill jigs 346, 493, 502, 733
End and face mills 326
Index 865

End clearance 272, 274 Face cutting-edge angle 318, 320


End clearance angle 273 Face mill 798
End cutting edge 274 Face milling 192, 312
End cutting-edge angle Face-milling cutters 186, 328, 253, 798
(ECEA) 187, 252, 275 Face-milling fixtures 546, 548
End-cutting machine reamer 351 Faceplate fixtures 565-567
End-milling cutters 798, 799 Faceplates 561
End mill (profile relief) 317 Face width 321
End mills 316, 797, 817 Facing 566, 809
End reamers 84, 85 Factors affecting forces and power 216
End reaming 83 Factors affecting heat treating 164
End relief 185, 187 Fatty emulsifiable mineral oils 266
End relief angle (ERA) 273, 275 Fatty mineral oils 265
Endurance limit 144 Fatty oils 263
Engine lathes 770 Feed 217, 250
Entrance angle 318, 730 Feed change gear plate 835
Epoxies 154, 730, 732 Feed change gears 834
Epoxy 731 Feed correction factors 202
Epoxy plastic 733 Feeding length 601
Epoxy plastic tools 731 Feeding speed 601
Epoxy resins 730, 733 Feed per cutter revolution 249
Epoxy tooling 731, 732 Feed per minute 249
Epoxy tooling materials 573 Feed per tooth 248, 323
Equalising rocker 442 Feeler gauges 101
Etching steel 117 Feeler-pin gauges 395
Expanding block-type boring tool 303 Feeler surfaces 544, 545
Expanding collet mandrel 570 Felt bobs 54
Expanding mandrels 569, 570 Ferrite 161
Expansion hand reamer 356 Ferrous tooling materials 145
Expansion machine reamer 356 Fibre-glass cloth 730, 733
Expansion reamers 355 File bands 38
Extended jaws 540 Files guides 99
Extension or pulley tap 799 Fillet 318
External broaching 559 Filling 98
External broaching fixture 560 Filling machines 36, 100
External center Internal center 362 Filllet 322
External (male) centre 361 Fine blanking 624
Extra-length drills 338 Finished design 5
Extra-thin-wall bushings 92 Finishing cut 252, 353
Extreme-pressure emulsifiable oils 266 Finishing hand taper reamer 356
Extreme pressure (EP) agents 708 Finishing operation 625
Extreme-pressure lubricant 705 Fixed gauges 391, 403
Extruding 620 Fixed limit gauges 392
F Fixed-program machine tools 769
Face 273 Fixed renewable 508
Face and shell mills 316 Fixed renewable bushings 93, 96
Face cutting edge 318 Fixture 434, 458
866 Index

Fixture design 434 Forming dies 52, 688, 739


Fixture design for numerically controlled Forming operations 700
machine tools 775 Forming tools 153, 841
Fixture jacks 441, 442 Form relief 315
Fixturing 776 Form-relieved milling cutter 314
Fixturing indicating gauges 421 Form tool 318
Flagging method 116 Four-flute core drill 339
Flagging technique 116 Four-flute hand taps 363
Flanged punches 630 Four-speed machines 833
Flange-forming 753 Free-machining additives 232
Flanges 709 French stop 661
Flank 272 Friction force 194
Flank wear 234 From-relieved cutter 315
Flash 697 Full shear 611
Flash change toolholder 803
G
Flat 344
Gall 695
Flat-bed press 670
Galling 689, 705
Flat clamp 96
Gap frame 594
Flat jig feed 507
Gap-frame presses 594, 601
Flat-leaf taper gauge 89
Gas-fired muffle furnance 167
Flatness 391, 415
Gas-fired semimuffle furnace 168
Flat relief 322
Gauge 390
Floating and adjustable holder 848
Gauge blocks 32, 399, 820
Floating reamer drivers 358, 359
Gauge tolerances 403
Flow-type air gauge 413
Gauge tools 735
Flushing 708
Gauging 390
Flush-mounted linears 94
Gauging methods 66
Flush-pin gauges 394, 395
Gear cutter (form relief) 317
Flute 318, 325
Gear cutting 267
Flute helix 324
Gear-driven presses 596
Flute length 321
General considerations in the design
Flutes 324, 340, 614, 619
of drill jigs 501
Flywheel 595
Geometry of single-point turning tools 277
Fool proofing 434
Glass cloth 735
Fool proofing pins 434
Glass-cloth ribbon 738
Forced vibrations 253, 254
Glass-laminate honeycomb structures 738
Forces in metal cutting 194
Glazing 351
Forged boring cutter 295
Go-not-go gauges 397, 398
Forging 157, 830
Grid pattern 779
Forging dies 52
Grid plate 780
Form 694
Grid system of fixturing 781
Formation of a chip 334
Grinding 68
Form cutters 314, 318
Grinding fixtures 571
Form die block 689
Gripper plate 623
Form gauges 397
Groove 290
Forming 598
Ground-in 290
Forming and drawing dies 153
Ground-in chip breakers 289
Index 867

Ground-in groove breaker 291 Hexagonal rest button 507


Ground-in groove-type chip breaker 291, 292 High-carbon-high-chromium cold-work
Group A 151 steels 152
Group D 152 High-carbon steels 161
Group F 153 High-helix drills 338, 339
Group H 152 High-speed boring 299
Group H11 to H16 152 High-speed drill press 101
Group H20 to H26 152 High-speed tool steels 152
Group H41 to H43 152 Hitch feeding mechanism 603
Group L 153 Hold-down clamps 544
Group O 151 Holding and locating the work 537
Group P Provide space Group P 153 Hole hog 84
Group S 151 Hole location 60
Group W 145 Hollow reamers 355
Guerin process 623 Honing 267
Guide 353 Hook 319
Guide bushings 554, 791 Hook angle 187
Guide pins 612 Horizontal and vertical boring operations 553
Guide posts 612 Horizontal rocking jaws 564
Guides 628 Hot-work tool steels 152
Guiding stock 658 Hydraulic and pneumatic clamping 465
Guid plate 700 Hydraulic power clamping 466
Gun or spiral-point tap 799 Hydraulic press 623, 624, 596
Gun tap 363 Hydraulic-press dies 732
Hydraulic-press forming 732
H
Hydraulioc indexing 547
Half-point angle 187
Hydrodynamic lubrication 263
Half-Side milling cutter (profile relief) 317
Hydropneumatic die cushions 692, 693
Hand-clamping fixtures 546
Hydropneumatic pistons 717
Hand finishing and polishing 43
Hand-reamer 351, 352 I
Hand scriber 61 Impingement 625
Hardenability 158 Inclination 281, 330
Hardening 68, 159, 730 Independent four-jaw chucks 561
Hardness 142 Indexable-insert face-milling cutter 331
Headless press-fit bushing 93, 508 Indexable-insert milling cutters 330, 332
Head press-fit bushing 508 Indexable-type tool 284
Heat treating 157 Indexing 805, 815, 816, 825
Heat treatment and tool design 173 Indexing fixtures 153, 546, 558
Heel 272 Indexing jigs 498
Heel drag 325, 326, 334 Index jig 498
Height gauge 61 Index milling 546
Helical 319 India oilstones 44
Helical flutes 318, 321, 353, 354 Indicating gauges 391, 407
Helical rake 319, 340, 341 Indicating parallel gauge 410
Helical teeth 321 Indicator depth gauge 408
Helix 362 Indicator-drop method 328
Helix angle 185, 187 Indicator plug gauge 409
868 Index

Indicator set 80 K
Indirect pilots 637 Kennametal de vibrator-k-bar 258, 259
Industrial materials 135 Kennametal roller de vibrator 255, 256
Information blocks 7 Keyway-broaching fixture 559
Inhomogeneous chip 189 Kicker 717
Injection or compression molding 153 Kirksite 52, 321, 733, 739, 741
Inner ram 718 Kirksite blanking dies 623
Insert 286 Kirksite cores 732
Inserted-tooth face-milling cutter 797 Kirksite dies 624
Insert indexing 287 K land 319
Insert punches 630 Knee tool 852
Inserts 284, 292 Knock collar 620
Insert-type milling cutters 797 Knock-off mandrel 570
Inside calipar 89 Knockout 595
Inspection fixtures 735 Knockout pin 617
Inspection gauges 390, 403 Knockout plate 620
Installation of drill bushings 90 Knockout rods 620, 690
Installation of renewable bushings 96 Knuckle-joint and hydraulic presses 697
Interlocking 319 Knurling swing tool 855
Interlocking milling cutters 314 Knurling tools 853, 855
Interlocking side 316 Kwick-switch tool holder 802
Interlocking-side milling cutter
L
(profile relief) 317
Lacquer 734
Internal (female) centre 362
Laminated tool 737
Inverted dies 617
Laminating 734, 735
Inverted die set 618
Laminating methods 736
Inverted single-action die 717
Lancing 592, 671
Ironing 683, 708
Land 319, 96
J Land width 353, 306
Jig 434, 443, 445, 448, 449, 458, 493, Lank wear 237
501, 510 Laps 57, 58
Jig borers 65, 66, 770 Lard oil 263, 265
Jig boring 64, 83 Latch clamps 462
Jig-boring machines 300 Lathe fixtures 561, 733
Jig-boring practise 75 Lathe-turning 198
Jig design 434 Layout 61
Jig feet 503, 506 Lead 319
Jig feet and legs 506 Lead angles 273, 332, 336
Jig grinder 65 Lead cam 839
Jig grinding 64 Leader pins 612
Jig leg 507 Leaf jigs 493, 494
Jig or a fixture design 435 Leaf-type diameter 494
Jigs and fixtures 153 Length gauges 393
Jobber’s-length drill 339 Length of shank 362
Jobber’s-length straight-shank drills 337 Length of square 362
Index 869

Length of stroke 596 Machine oils 263


Line boring bars 297-299 Machine tool 182
Line boring operation 554 Machine-tool efficiencies 203
Liner bushings 507 Machining 82
Lip 320, 451 Machining centre 770, 771, 807, 813, 815
Lip angle 320 Machining fixtures 733
Lip relief 187, 436 Magnesium 157
Lip relief angle 335, 438 Magnetic and vacuum chucks 552
Lips 336 Magnetic chuck 550
Liquid-bath furnaces 167 Magnetic-chuck fixtures 573
Liquid-bath, or pot, furnaces 168 Magnetic chucking devices 572
Liquid laminating plastic 735 Magnetic-chuck parallels 572
Locating 434 Magnetic chucks 549
Locating an edge 82 Magnetic-chuck V blocks 572
Locating devices 451 Magnetic clamping 549
Locating from circular surfaces 438 Magnetic cylinder squares 30
Locating from irregular surfaces 441 Magnetic fixture 540
Locating from plane surfaces 435 Magnetic lathe chuck 568
Locating methods 435, 441, 451 Major burr 501
Locating methods and devices 444 Mandrel 570
Locating microscope 35 Mandrel for threaded parts 571
Locating pin 445 Mandrels 561, 568
Locating points 438 Manual feeding 599
Locating surface 435 Margin 335
Location 437 Martensite 164
Location and clamping 501 Masonite 154
Location errors 439 Mass casting 734, 742
Location of blanks 663 Mass-cast metal-forming tool 743
Location of rough forging 444 Master bushings 509
locators 67, 437 Master die 114, 324
Lock screw 93 Master gauges 392, 825
Lock-screw locating pins 446 Master mould or model 735
Lock screws 94 Master patterns 114, 118
Long-angle 344 Master setting gauge 824
Longitudinal thrust force 198 Matensite 164
Low-alloy special-purpose tool steels 153 Material-handling equipment 598
Low-carbon mold steels 153 Measurement 89
Low-carbon steels 161 Mechanical 290
Low-helix drills 339, 338 Mechanical clamps 552
Low-melting bismuth alloys 128, 129 Mechanical dynamometer 199
Low-melting tooling materials 125 Mechanical equalising jacks 442
LP relief angle 342 Mechanical indicators 407
Lubricants 263 Mechanical or clamped-on
Lucite 124 chip breaker 291, 292
Mechanical press drives 597
M
Mechanical presses 596, 623
Machinability 143, 221, 230-232, 238, 239
Mechanics and geometry of
Machinability ratings 221, 222, 224, 226, 230 chip formation 187
870 Index

Melt-away tooling 126 Neck 321


Merchant’s circle diagram 194, 751 Needle files 45
Metal cutting 182, 183, 552 Needs analysis 3
Metal-cutting band saw 40 Negative radial rake 321
Metal-cutting tools 271 Negative rake 319
Metal forming 739, 749 Nest 501
Metal-forming dies 731 Nesting 319
Metal-slitting saw (profile relief) 316, 317 Nest location 450
Method of clamping 452 Nest or cavity location 449
Methods of breaking chips 345 Nest or locating surfaces 452
Methods of conical location 439 Neutral axis 685
Methods of V location 440 Neutral plane 685
Micrometer 61 Nomenclature of reamer 352
Microsine plate 79 Nomenclature of tap 361
Microstructure 160 Nomenclature of twist-drill 335
Mild or low-carbon steel 153 Nonmetallic tooling materials 154, 155
Mill fixture 542, 775 Nonsticking plugs and locators 447
Mill-fixture base 542 Normal force 194
Mill fixtures 545, 557 Normalizing 157
Milling 267, 545, 806 Normal rake angle 340
Milling cutters 307, 311, 314, 323, 733, 796, Nose 273
797 Nose radii 218
Milling fixtures 541, 544, 549 Nose radius 220, 272, 594
Milling machines 65, 118, 769 Nose-radius recommendations 278
Milling-machine vise 778 Notching 593
Milling operations 328, 546 Novaculite 44
Mineral-based cutting oils 264 Numerical contro 769
Minor burr 501 Numerical Control 770
Misfeed detectors 638 Numerically controlled drilling machine 771
Mist cooling 271 Numerically controlled machining
Mist generator 270 centre 772, 773
Mockups and moulds 733 Numerically controlled turret drill 806
Modular fixture 783
O
Modulus of elasticity 137
OBI press 594, 595, 692, 693
Molybdenum hot-work steels 152
Oblique cutting 192
Motor-driven coil-unwinding reel 599
Offset 322
Mould or model preparation 734
Oil emulsions 263
Mounted point 50
Oil-hardening cold-work steels 151
Mounted wheels 50
Oil-hard steels 158
Muffle type 167
Oil-hole drills 268
Multiple-cutter boring 299
Oil quenching 172
Multiple-diameter drills 339
On-end or slug-type tools 284
Multiple gear cutter (form relief) 317
Open jig 496
N Optical comparator 417, 823
National coarse (NC) 396 Optical flats 403
National fine (NF) 396 Optical projectors 417
N/C drilling machine 770 Optical tool setter 822
Index 871

Optical tool-setting gauge 822 Piloted reamers 358


Or crankshaft 344 Pilot pin 669
Orthogonal and oblique cutting 191, 198 Pilots 353, 628, 636
Orthogonal cutting 192, 193 Pilot types 636
Other grinding 267 Pin and button locators 444
Outer ram 718 Pinch clamp 454, 455
Overbending 683 Pinch clamp with wedge 463
Overforming 753 Pinch trim 718
Oxide cutting tool materials 154 Pin gauges 393, 394
Pin or stud fixtures 546
P
Pins and buttons as locators 445
Pack hardening 159
Pin-type clamping mechanism 287
Pad-type form dies 689
Pitman 595
Pantographs 733
Plain 316
Pantomills 123, 415
Plain and side mills 316
Paper-base phenolic sheet stock 738
Plain blanking 592
Parallel 290, 61
Plain hollow mills 845-847
Parallelism 391, 415
Plain mandrels 568, 570
Parallel setup block 76
Plain milling cutter (profile relief) 317
Parting 593
Plain milling cutters 325
Paste construction 734
Plain mills 324
Paste plastic 747
Plain plug gauge 396
Paste resins 744, 745
Plain punches 627, 629, 630
Pawl stop 659
Plain screw bushing 510
Pearlite 161
Plain-wedge clamp 463
Pedestal punches 627, 630
Planer gauge 32, 741
Pedestal-type punch 630
Planers 243
Pencil trace 122
Planing 272
Penetration 605
Plastic 741
Percent reduction 709
Plastic dies 744
Perforator 636
Plastic draw and form dies 732
Perforator (insert) punches 631
Plasticity 136
Perforator punches 627, 632
Plastic masters 735
Perforator-type punches 631
Plastic patterns 739
Peripheral cutting edge 320
Plastics 154, 729, 730
Peripheral milling 310
Plastic tool 742
Peripheral relief 187
Plastic tooling 154, 734
Perpendicularity (squareness) 391
Plastic tooling methods 734
Phenolics 154, 730
Plastic tubes 738
Pick feed 119
Plastic-tubing applications 738
Pick-feed duplicating 121
Plate jigs 496, 497
Pickling 709
Plates 503
Pierced 607
Plug and ring thread gauges 395
Piercing 332, 415, 591, 592
Plug-gauge fit 85
Piercing die 755
Plug gauges 89, 393
Piercing point 655
Plugs 444
Piercing punch 618
Plug tap 363
Pilot dimensions 636
872 Index

Pneumatic air cushions 692 Press feeding 598


Pneumatic gauges 412 Press-fit and linear bushings 93
Pneumatic gauging 412 Press-fit bushings 507
Pneumatic-type hand grinders 46 Press-fit pilots 637
Pocket-size tap-drill 367 Press fitting 699
Point 336 Press-fit wearing bushing 508
Point angle 335 Press-lubricant 706
Point diam 362 Press speed 601
Point diameter 362 Press tonnage 598
Point grind 342 Press tools 67
Pointing toolhoder for circular tools 853 Pressure-fed mist generator 271
Pointing tools 852, 853 Pressure-pad forming dies 691
Polar co-ordinates 74 Pressure pads 638
Polycheck 789 Pressure pin 693
Polyesters 154, 730 Pressure plate 115
Polymerising 730 Pressure-pot method 741
Positioning system 770 Pressure-pot system of surface casting 742
Positive and negative disposable inserts 285 Pressworking lubricants 708
Positive rake 315, 319 Primary die stop 618
Positive rake angles 276 Primary relief 322
Positive rake tools 284 Principles of Location 434
Positive stop screw for turret slide 844 Profile grinder 42
Pour method 740 Profile relief 316
Power-clamping fixtures 546 Progressive dies 617, 619, 689
Power hand grinders 46 Progressive fixtures 547
Power hand-polishing equipment 46 Progressive form die 690
Power-press types 594 Progressive milling 546
Power required for drilling 349 Properties of materials 135
Power requirement for drilling 346 Punch 63
Precesion drill chucks 83 Punch and die manufacture 95
Precheating 160 Punch and die mountings 612
Precision die sets 614 Punch design 627
Precision presetting 821 Punched tape 770
Presetting 816 Punches 102
Presetting fixtures 819 Punches dies 628
Presetting height gauge 819 Punches in punch plates 633
Presetting radial-cutting tools 819 Punches mounted in punch plates 627, 632
Presetting tools 821, 822 Punch height 630
Presetting tools for N/C lathes 822 Punch holders 612, 620
Preset tools 816 Punching 666
Press-brake 752 Punching pressure 603
Press-brake bending and forming dies 682 Punch-noise radius 710
Pressed-in 290 Punch-nose 712, 705
Pressed-in chip breaker 293 Punch-nose radius 709
Pressed-in, or molded 293 Punch pilots 618, 636
Pressed-in, or molded chip breaker 292 Punch plates 618, 632
Presses 594 Punch presses 662
Index 873

Punch radius 702 Reciprocal milling 546


Punch shedders 635 Reciprocating filling machine 37
Punch shoe 63 Recommended widths of chip breakers 291
Push toggle clamp 464 Rectangular box-shaped cups 712
Redrawing 598
Q
Redundant location 435
Quenching 159, 163
Reed magnifier 411
Quenching process 172
Reed mechanism 411
Quenching stresses 174
Reference gauges 390
Quick-acting 456
Regular mineral-oil emulsion 265
Quick-acting clamps 546, 698
Relief 325
Quick-acting screw clamp 460
Relief angles 184, 322, 325, 307, 503
Quick-change toolholder 802-804
Relieved 319
Quick-change tooling 800, 804, 805
Relieved land 322
Quick-change tool post 805
Renewable bushings 94, 508
R Research and Ideation 4
Rack-and-pinion clamps 465 Rest blocks 543
Radial 362 Rest button 507
Radial-cutter boring bar 300 Rest pads 449
Radial-cutter boring head 299 Rest pads and plates 448
Radial drill presses 300 Rest pins 441
Radial force 198 Rest plates 449
Radial rake (–) 186, 187, 306, 330 Retaining ring 696
Radial rake angle 320, 322, 323 Rib-and-cavity embossing die 756
Radial relief 101, 322 Rib-embossing die 756
Radial relief angle 320 Right- and left-hand knee tool 852
Radial run out 319 Rigidax 126
Radius-forming 753 Rigidity 501
Radius-forming die 752 Ring gauges 392
Radius on punch 703 Ring thread gauge 396
Railroad wheel-tread form gauge 397 Rise-and-fall fixtures 546
Rake 192, 319 Riveting 699, 700
Rake and relief angkes for procision-boring Roller-backrest box tool 847
tools 308
Roller-coating 708
Rake angles 184, 186, 288, 355, 361
Roller stock pusher 639
Ram 595
Roll feed 601
Rapid pull-out attachment 838
Roll forming with urethane 758
Rapid-pullout mechanism 844
Rolling 695
Rapid tooling 666
Rolling mills 680
Reamer 299, 302, 393, 848
Rollover 603
Reamer classification 354
Root 362
Reamers 326, 346, 354, 360, 805
Rose or fluted reamers 85, 86
Reaming 243, 323, 328, 346, 353
Rose reamers 351
Reaming operation 91
Rotary cross-slide milling-head
Reaming process 346 attachment 100
Rear upper slide 845 Rotary cutters 49
Rear-upper slide adjustment 846 Rotary fixtures 49, 100, 546, 549
Receiving gauges 392 Rotary milling 546
874 Index

Roto recipro toolmaking machine 102 Series lever magnifier 408


Roughing cuts 252 Set back 685
Roughing hand taper reamer 356 Setup blocks 76
Rough milling 102 Seven-station rotary fixture 549
Round end clamp 96 Shaft-eccentric cam clamp 458
Round pin or button 444 Shaft eccentric clamp 459
Rubber 154, 697 Shaft template gauge 397
Rubber flex collets 561 Shank 272, 353, 594
Rubberised abrasive points 51 Shank diam 362
Rubber-pad blanking 623 Shank diameter 335
Shank length 335
S
Shape factor 761
Sanding 732
Shapers 243
Sawing 267
Shaping 272
Saws 316
Shaving 101, 593
Scaling 169, 281
Shear angle 188
Scanning 119
Shearing 137
SCEA 275
Shearing force 194
Scrap-web allowance between 666
Shear plane 188
Scrap-web allowances 665
Shear strain in metal cutting 197
Screw bushing not guided by thread 510
Shear strength 136
Screw bushings 510
Shear stress in metal cutting 197
Screw clamps 459
Shedder pin 636
Screw holes 56
Shedders used on punches 635
Screw-machine-length drill 339
Sheet-metal drilling 337
Screw-on boring cartridge 304
Sheet-metal presswork 591
Screws 55, 629, 689
Shell end mill (profile relief) 317
Screws and dowels 625
Shelling off 165
Screw-slotting 316
Shell reamer and arbor 356
Screw-slotting cutter (profile relief) 317
Shell reamers 355
Screw-type clamping mechanism 286
Shims 558
Scriber set 61
Shim stock 63, 101
Scribing 100
Shock-resisting tool steels 151
Secondary relief 323, 326
Shop gauges 390
Secondary shear 605
Short-run tooling 666
Secondary standard 400
Shoulder stops 659
Secondary trimming 625
Shut height 596, 598
Sectional components 628
Shuttle broaching fixtures 560
Sectional die blocks 627
Side 316
Sectioning large die 628
Side clearance 272
Segmental 503
Side clearance angle 273
Segmental chips 189, 503
Side cutting-edge 187, 274
Seizure 705
Side cutting-edge angles 219, 272, 306
Selection of stripper strings 643
Side knurl holder 855
Self-excited 254
Side-loaded jigs 92
Self-locking latch 462
Side milling cutter (profile relief) 317
Semimuffle furnace 168
Side of thread 363
Semimuffle type 167
Index 875

Side-pressure forming 753 Slugs 603, 608


Side rake 187, 274, 353 Snap gauges 393, 394
Side-rake angles 273, 278, 306 Socket-head cap screws 56, 506, 511, 626
Side relief 187 Solid-carbide end mills 798
Side relief angle (SRA) 273, 275 Solid form dies 689, 690
Sight locators 443 Solid jaws 564
Signature 274 Solid reamer 355
Silicon carbide 44 Spade drills 796, 801
Silicon carbide abrasive 58 Special bar tools 304
Silver brazing 283 Special jaws 540
Simple dies 67 Special-purpose tool steels 153
Simple lever magnifier 407 Special top jaws 562
Sine bars 31 Special vise jaws 540
Sine plates 32 Specific cutting speed 226
Single-action draw die 717 Specific energy 200, 220
Single-action press 597 Speeds and feeds for high-speed-steel
Single- and double-action draw dies 716 reamers 247
Single-angle milling cutter (profile relief) 317 Speeds and feeds for shapers and
Single-piece milling 546 planers 248
Single-piloted bar 552 Speeds and feeds used for drilling 245
Single-piloted boring bar 554 Speeds and lubricant for machine
tapping 367
Single-point boring 346
Sperm oil 263
Single-point boring machine 555
Spherical locating plugs 448
Single-point boring tools 632
Spherical washers 453, 454
Single-point cutting tools 271
Spheroidise annealing 232
Single-point metal-cutting tools 272
Spheroidising 157, 232
Single-point turning-tool geometry 276
Spindle 832
Single-ram vertical-surface broaching
machine 560 Spindle and feed adjustments 838
Single-roll feed mounted on an OBI press 601 Spindle-mounted milling cutters 313
Single shear on die 611 Spindle speeds 832, 833
Single shear on punch 611 Spiral-cam edge clamp 458, 459
Sintered carbides 155 Spiral cutting of round plastic tubing 737
Slab-milling fixtures 546 Spiral drills 501
Sleeve-type expanding mandrel 569 Spiral flute 353
slender punches 634 Spiral-fluted hand taps 364
Slide (or hitch) feed 601 Spiral-fluted tap 364, 799
Slide or hitch feeding mechanism 602 Spiral flutes 355
Slide or hitch feeds 602 Spiral point 793
Slide tool 849 Spiral-point drill 792
Slip 188 Spiral-point hand taps 363, 364
Slip planes 188 Spiral reamers 355
Slip renewable bushings 96 Spiropoint drill sharpener 793
Slitting saws 324 Spline tools 748
Slotting fixtures 546 Split 344
Slug ejection 625 Split-point drill 344
Slug pulling 635 Split sleeves 570
Splitting 63, 339, 344
876 Index

Splitting the point 341 Standard vises and chucks 778


Spotters 84 Starting taper 353
Spotters drills 85 Statement of the problem 3
Spotting drill 793, 794 Stationary fixture 556
Spotting-or centre-drill 794 Steel-cutting grades 374
Spotting tool 83 Steel-rule dies 621, 622
Spraying 708 Step drills 340
Spring 640 Step-head punch 632
Spring-actuated shedder pins 635 Stepping punches 611
Springback 608 Stock 592, 638
Spring collet and stock-feed mechanism 835 Stock guide 689
Spring collets 561 Stock stops 656
Spring-loaded knockouts 717 Stock straighteners 600
Spring-loaded pilots 637, 638 Stock-strip stops 659
Spring-loaded shedder pin 636 Straddle-milling fixture 546
Spring-loaded stripper 609 Straight cutting oil 264
Spring-operated strippers 638, 639 Straightedge 61
Spring plates 619 Straight emulsifiable mineral oils 266
Spring strippers 642, 658, 660 Straight flutes 355
Spring-type pressure pads 692 Straight grades 368
Square 363 Straightness 415
Squared shank 354 Straight shank 335
Squareness 415 Straight-shank drills 337, 339
Squareness-indicating gauge 410 Straight-shank extra-length drill 339
Square tools 851 Strap clamps 453, 777
Squash method 741 Strap clamps and heel blocks 775, 776
Stabilising 158 Strap clamps employing the wedge
Staggered-tooth milling cutter principle 463
(profile relief) 317 Strength 135
Staggered-tooth side 316 Strength of materials 137
Staking 699 Stress relieving 158
Stamping 706 Stretch dies 732
Standard 344 Stretch-form dies 730
Standard buttons and pins 445 Stretching 685
Standard clamps 452 String milling 546
Standard die clearances 608 Strip 620
Standard die set 613 Strip-feeding equipment 601
Standard edges for stampings 606 Strip layout 663, 664
Standard hand taps 363 Stripper 659
Standardized automatic stop 661 Stripper 618
Standard jaws 562 Stripper bolt 640
Standard mandrels 561 Stripper-opening clearance 639
Standard milling-machine vise 779 Stripper pins 690, 691
Standard mill vise 790 Stripper plates 102, 639
Standard pilots and punches 637 Strippers 622, 638
Standard strap clamp 453 Stripper springs 620, 655
Standard vertical turret tool holders 809 Stripping 609, 692
Index 877

Stripping devices 609 Taper reamers 357


Stripping edge 700 Taper shank 323
Stripping force 644 Taper-shank drills 338, 339
Strip-straightening devices 600 Taper tap 363
Stroke 598 Tap holders and taps 849
Stub and line boring bars 302 Tapped 632
Stub bars 296 Tapping 56, 364
Stub boring 295 Tapping head 800
Stub boring bars 297 Tapping operations 363
Stub screw-machine reamers 355, 356 Tapping problems 365
Stylus 102 Tapping process 360
Subland drill 339 Tapping speeds 367
Sulphochlorinated mineral oils 265 Taps 817
Sulphurised fatty mineral oils 265 Taps 251, 276, 360
Sulphurised mineral oils 265 T bolts 544
Surface casting 734, 739 Tearing 691
Surface-casting plastics 740 Telescoping gauge 89
Surface finish 252 Temering 159
Surface gauge 61 Template gauge for machine ways 397
Surface-gauge scriber 61 Template gauges 397
Surface roughness 252 Template jigs 497, 496
Surface texture 391 Temporary fixture 548
Surge tank 692 Temporary indexing fixture 547
Swabbing 708 Temporary mill fixtures 547
Sweeping tools 84 Tensile strength 135
Swing-clamp assembly 461 Tension 137
Swinging-index-type fixture 567 Tentative design solutions 5
Swinging-index types 566 The basic principles of clamping 452
Swinging-type strap clamp 456 Thickness 415
Swing tool 849 Thread 363
Swivel block 87 Thread-cutting oil 263
Symbolic control 769 Thread gauges 396
Thread grinding 267
T
Threading 267, 363
Tab stake 700
Thread plug gauge 396
Tang 335
Thread relief (radial) 363
Tang drive 335
Thread relief radial 362
Tangential force 198
Thread ring gauge 396
Tangential rake 322
Thread rolling 267
Tantalum carbide 374
Threads 56
Tap 56, 800
Three- and four-flute core drills 338
Tap Classification 363
Three-jaw chuck 779
Tap drill 56
Three-jaw universal chucks 561
Tap driver 800
Three-wire system of thread
Tapered cutters 123
measurement 419
Taper gauges 399
Throwaway insert-type toolholder 801
Taper-length straight-shank drill 339
Throwaway or disposable insert 284
878 Index

Thrust in kN for drilling 347 Top jaws 561


Thumbscrew latch 463 Top knurl holder 855
Toe clamps 454 Top turret slide 844
Toggle clamps 464 Total indicator variation 354
Toggle fixtures 546 Toughness 136
Tongs 173 Tracer and cutters 122
Tool 145, 322 Tracer and duplicating mills 118, 123
Tool and diemaker 731 Tracing equipment 119
Tool carrier 816 Transfer 62, 63
Tool changers 772, 813, 805 Transfer methods 64
Tool character 274 Transfer punch 63, 64
Tool-chip interface 233, 729 Transfer screws 63
Tool-chip interface temperature 237 Transfer screws and punches 63
Tool codes 811, 812 Transfer tools 29
Tool designer’s 182, 187, 444, 731, 830 Transformation range 162
Tool failures 165, 368 Transition error 64
Tool form gauge 397 Trim allowance 715
Tool geometry 185, 187 Trimming 594, 620
Tooling 1, 669, 830 Trip dogs 119, 840, 842
Tooling for automatic screw machines 845 Triple-action press 597
Tooling materials 135, 731 Trip lever 839
Tooling plastic 741 Trip-lever abutments 840
Tool life 238 Trip levers 836
Toolmakers 449, 666 Trip stop 658, 659
Toolmaker’s block and clamp 23 Troubleshooting reaming problems 357
Toolmaker’s buttons 29, 62 Trowelling 731
Toolmaker’s clamps 30 True rake 281, 320
Toolmaker’s dial test indicator 25 True-rake angle 278
Toolmaker’s hammer 24 T slot 544
Toolmaker’s lathe 23 T-slot cutter (profile relief) 317
Toolmaker’s machines 23 T slots 542
Tool offsets 822 T-slotted bolster plate 670
Tool overhang 255 Tungsten carbide 368
Tool positioners 805 Tungsten carbide-cobalt 374
Tool presetter 822 Tungsten hot-work steels 152
Tool Presetting 816-818 Turning 192
Tool-presetting fixture 823 Turning swing tool 849
Tool setter 816 Turret 841
Tool-setting 817, 821 Turret adjustable knurl holder 853
Tool-setting gauge 824 Turret dog 842
Tool shank 283 Turret drilling machine 805
Tool signature 274, 306 Turret indexing 806, 842
Tool Steels 145 Turret lathes 770
Tool temperatures 237 Turret slide 842, 844
Tool wear 104, 234 Turret-slide drive 844
Tooth 320 Turret-slide rack 843
Tooth face 320 Turret-type N/C machines 805
Index 879

Twist drills 185, 333, 337 Urethane draw dies 757


Twisted-strip mechanical indicator 411 Urethane elastomers 749, 751
Two-bladed boring bar 303 Urethane embossing dies 755, 757
Two-flute drill 500 Urethane fixture jaws 759
Two-way magnetic chuck 550 Urethane fixtures 760
Two-way magnetic clamping 550 Urethane forming die 752
Types of boring tools 295 Urethane inside-flanging dies 758
Types of clamps 452 Urethane pressure pads 753, 754, 760
Types of drill points 344 Urethane punch 757
Types of drills 337 Urethanes 154, 655, 731, 760
Types of fixtures 539 Urethane strippers 657
Types of mill fixtures 545 Urethane stripping pad 755
Types of single-point tools 282 Urethane strips 754
Typical strap clamp 454 Urethane tooling 752
Urethane wear pads 759
U
Urethane wipe-down dies 752
U bending 687
Urethene 656
U dies and channel dies 686
Urethene inside flanging die 758
U-forming 753
U-forming die 752 V
Uneven quenching 175 Vacuum chuck 551
Universal fixtures 547, 558, 775, 783, 785, Vacuum chucking 549
786 Vacuum chucks 550
Universal jigs 458, 465 Vacuum-contoured chuck 568
Universal or pump jigs 499 V-bend die 682
Universal pedestal base fixture 785 V-bending 681, 683, 686
Universal punch shaper 42 V-block fixtures 546
Universal self-centreing scroll chucks 561 V-die air bends 683
Universal vacuum chuck 551 Velocity relationships 193
Universal vertical fixture 786 Vertical milling machines 100, 300
Universal vertical vise 790 Vertical rocking jaws 565
Unpowered plain coil-unwinding reel 599 Vertical-spindle miling machines 66
Upper die 696 Vertical toggle clamp 464
Upper slide adjustments 838 V-forming 752, 753
Upper-slide cams 839 Vibration damping 257
Upset stake 700 View layout 11
Upside-down EDM maching 112 Vise 546
Upside-down machining 111 Vise fixtures 539
Urethane adhesive 752 Vise jig and cutter-setting gauge 549
Urethane bulging 698 Vise jigs 499
Urethane bulging dies 755 Visual thread gauge 421
Urethane chuck jaws 759 V locating methods 441
Urethane clamping jaws 759 V method 439
Urethane compounds 731
W
Urethane die 750
Water-hardening tool steels 145
Urethane-die designs 752
Wear land 322, 323
Urethane-die mechanism 749
Wear strips 156
Urethane dies 749, 759
Web 335, 337, 340, 345
880 Index

Web thickness 337 Woodruff key-seat cutter (profile relief) 317


Wedge clamp 462 Woodworth tables 74
Wedge-type holder 303 Workhardened 250
Wedge-type jaws 540 Work hardening 232
Welded frames 594 Work-holding 549
Weld fixtures 537 Work-holding fixtures 561
Welding 705, 708 Work horn 558
Whistler magna-die 668 Working stroke 592
White rehardened zone 106 Wraparound jaws 565
Width gauges 393 Wrenchless adapter 804
Width of land 326 Wrinkles 702
Wipe-down die 753 Wrinkling 704
Wiper blade 332 Wrung 403
Wiper nest 332 Z
Wiping bend 687 Zero clearance 504
Wiping dies 681, 683, 684 Zero rake 315
Woodruff key 316 Zinc-base alloys 156

You might also like