Sie sind auf Seite 1von 432

Engineering Plasticity

Engineeri ng Plasticity
Theory and Application to Metal
Forming Processes

R. A. C. Slater
Department of Mechanical Engineering
The City University, London

M
© R. A. C. Slater 1977

Softcover reprint of the hardcover 1st edition 1977 978-0-333-15709-1

All rights reserved. No part of this publication may be


reproduced or transmitted, in any form or by any means, without permission

First published 1977 by


THE MACMILLAN PRESS LTD
London and Basingstoke
Associated companies in New York Dublin
Melbourne Johannesburg and Madras

ISBN 978-1-349-02162-8 ISBN 978-1-349-02160-4 (eBook)


DOI 10.1007/978-1-349-02160-4

Photosetting by Thomson Press (India) Limited, New Delhi

This book is sold subject to the standard


conditions of the Net Book Agreement
Contents

Preface viii

1. Introduction 1
1.1 Definition and Scope of the Subject 1
1.2 Nature of Engineering Plasticity 2
1.3 A Brief Historical Account 3
1.4 Classification of Metal Forming Processes 5
1.5 Forming Limits 7

2. Stress Analysis 11
2.1 Specification of Stress at a Point 11
2.2 Differential Equations of Equilibrium in the Neighbour-
hood of a Point 16
2.3 Three-Dimensional Stress Analysis 20

3. Strain Analysis 37
3.1 Infinitesimal Deformation 37
3.2 Finite Deformation 51

4. Yield Criteria for Ductile Metals 67


4.1 General Considerations 67
4.2 Von Mises Yield Criterion 69
4.3 Tresca Yield Criterion 70
4.4 Yield Surface for an Isotropic Perfectly Plastic
Material 71
4.5 Haigh-Westergaard Stress Space Representation of
Yield Criteria 73
4.6 Experimental Verification of Yield Criteria 79
4. 7 Subsequent Yield Surfaces for an Isotropic Strain-
Hardening Material 84
4.8 A Yield Criterion for an Anisotropic Material 86
vi CONTENTS

5. Stress-Strain Relations 90
5.1 Elastic Stress-Strain Relations 91
5.2 Elastic Strain Energy Functions 93
5.3 Plastic Stress-Strain Relations 95
5.4 Stress-Strain-Rate Equations 100
5.5 Plastic Work and Strain-Hardening Hypotheses 100
5.6 Experimental Verification of the Prandtl-Reuss
Equations 103
5.7 Derivation of the Generalised Plastic Stress-Strain
Relations 105
5.8 The Principle of Maximum Work Dissipation 115

6. Phenomenological Nature of Engineering Metals 119


6.1 Changes in Density and Shape 119
6.2 Quasi-Static Uniaxial Tensile Test 120
6.3 Quasi-Static Compression Tests 134
6.4 Effect of Temperature and Strain-Rate 141
6.5 Anisotropy and Aelotropy 152
6.6 Bauschinger Effect 157
6.7 Plastic Instability 159
6.8 Homogeneous Deformation 168

7. Plane Strain Plastic Deformation 175


7.1 Basic Concepts and Assumptions 175
7.2 Fundamental Equations for Plane Plastic Flow 176
7.3 Plane Strain Slip Line Field Theory 182
7.4 Some Applications of the Slip Line Field Theory 209
7.5 Estimation of Plane Strain Forming Parameters by the
Stress Evaluation Approach 236

8. Plastic Strain with Axial Symmetry 263


8.1 Fundamental Equations for Plastic Strain with
Axial Symmetry 263
8.2 Approximate Solution for the Quasi-Static Axisymmetric
Upsetting of a Circular Cylinder 266
8.3 Fast Homogeneous Compression of a Circular Cylinder 271
8.4 Bar and Wire Drawing Through a Conical Die 274
8.5 Extrusion Through a Conical Die 280
8.6 Optimum Die Angles 291
8.7 Hydrostatic Extrusion 292
8.8 Drawing and Sinking of a Thin-Walled Tube Through
a Conical Die 294
CONTENTS vii
9. Plane Plastic Stress and Pseudo Plane Stress 310
9.1 Basic Concepts and Assumptions 310
9.2 Equations of Plane Stress with von Mises Yield Criterion 312
9.3 Simple Stress States 316
9.4 Equilibrium of a Thin Plate with a Circular Hole Subjected
to Uniform Pressure 318
9.5 Plastic Bending of a Thin Circular Plate 321
9.6 Deep Drawing of a Circular Sheet Metal Blank 328

10. Extremum Principles for a Rigid-Perfectly Plastic Material 348


10.1 General Considerations 348
10.2 Basic Energy Equation or Virtual Work Equation 349
10.3 Surfaces of Stress and Velocity Discontinuity 351
10.4 Uniqueness Theorem for a Rigid-Perfectly Plastic
Body 354
10.5 The Lower Bound Theorem 356
10.6 The Upper Bound Theorem 358
10.7 Some Applications of the Upper Bound Theorem to
Plane Strain Deformation (Johnson Upper Bound
Estimates) 362
10.8 Upper Bound Estimates for Anisotropic Metals 382
10.9 Lines of Thermal Discontinuity or 'Heat Lines' and
Temperature Jumps During a Fast Metal Forming
Process 383
10.10 Plastic Bending of a Notched Bar-the Plastic Hinge 390
10.11 Upper Bound Estimates Involving Unit Deforming
Regions (Kudo Upper Bound Estimates) 392
10.12 Upper Bound Estimates for Axisymmetric Deformation
(Kobayashi Upper Bound Estimates) 398
10.13 Upper Bound Estimates for the Plastic Bending of
Transversely Loaded Thin Plates 401

Appendices 410
1 The Rule of Sarrus for the expansion of third-order
determinants 410
2 The characteristics of partial differential equations 412

Author Index 416

Subjectlndex 419
Preface

Except for castings, which are formed from the liquid state, all metal products
are subjected to at least one metal forming process during their manufacture.
In such processes the desired change in shape of the metal is effected without
metal removal. The deformation is permanent and involves predominantly
large plastic strains. The elastic strains are then of lesser significance.
With the current rapid technological progress, the theory of plasticity
has been brought forcibly into the forefront of engineering application
and design. The facts of economic life have made the efficient utilisation
of material, labour and time and a more efficient approach to design a
necessity even for the less sophisticated industrial applications. Today there
is, therefore, a greater need for all mechanical and manufacturing engineers
to be familiar with the fundamentals of metal forming analysis, not only to
be more numerate and hence reduce the extent of empirical approach
to problems, but also to close the gap which exists at the interface between
design and manufacture.
Although the theory of plasticity has advanced considerably in the last
few decades, rigorous solutions to all metal forming problems are still not
available. However, simplified versions of practical processes assuming
idealised materials and boundary conditions give valuable insight into the
effects of various parameters such as the extent of deformation, tool profile,
speed and lubrication on the force, energy and power requirements. The
resulting temperature changes, forming limits and the final properties of the
manufactured products can be assessed.
This particular text is based on lectures relating to the theory of plasticity
and its specific application to metal forming processes which have been
given during my university career at one time or another. They have been
addressed to advanced undergraduates commencing in the second year
of their studies and also to postgraduate students in mechanical engineering
and manufacturing engineering.
My principal aim in producing this text is, therefore, to bring within
the scope of advanced undergraduates and those studying for higher degrees
PREFACE ix
in mechanical engineering and manufacturing engineering, at universities
and polytechnics, a compilation of information and some analytical methods
in the subject of ENGINEERING PLASTICITY and its specific application
to metal forming processes. In addition, it is hoped that the account presented
of the principles of metal forming theory and application will be ofbenefit to
all practising engineers and metallurgists who are directly concerned with
improving design and efficiency in the manufacturing industry.
In the preparation of this book it was obviously impossible to include
all aspects of the metal forming processes. Rather than attempt to cover the
whole field in a cursory manner I have preferred to consider a wide range
of selected topics and present the treatment of these as complete and factual
as possible. All the theory of plasticity necessary for this purpose is included.
The textbook is not intended to be a mathematical treatise but, equally,
it contains little descriptive detail of the many complexities of modem metal
forming processes. For these reasons a mainly technological approach to
the subject is adopted and it is hoped that this is reflected in the choice
of title.
Metal forming, like all engineering activities, is essentially interdisciplinary.
It is therefore necessary, for example, for readers to be familiar with the
relevant metallurgical aspects. The various implications of physical metal-
lurgy or materials science which affect the applications of plasticity should
be borne in mind. However, it has proved impossible to include such infor-
mation in the present text but excellent textbooks are now available on this
subject.
It is often assumed that readers who have an interest in plasticity theory
and its various applications are familiar with the basic concepts of stress
and strain analyses. Nevertheless, it is my experience that this is not always
the case. Chapter 2 (Stress Analysis) and chapter 3 (Strain Analysis) are
consequently presented in some detail so as to avoid the need to refer to
other texts and also to refresh the memory of those readers who have,
indeed, previously studied these concepts.
Stress and strain as well as being vector quantities are also second-order
tensors. Although there is no intention to be mainly concerned with tensors
and their properties as such, it is important that readers are at least familiar
with the subscript notation known as 'tensor' notation. This notation is
not only a convenient and shorthand way of stating mathematical expres-
sions, but is also extremely useful in derivations and in the proof of some
theorems. Many research papers employ this type of notation. It is assumed
that the reader has acquired sufficient knowledge to interpret this notation
without much difficulty.
For the development of a plasticity theory, the stress states at which
plastic flow or yielding will occur, and the flow rules or relations between
stress and strain when plastic flow does occur, are necessary. Chapter 4
discusses yield criteria and their experimental verification but is restricted
X ENGINEERING PLASTICITY

to ductile metals which are of greatest importance in metal forming. In


chapter 5, the stress-strain relations are developed and discussed.
Having established the fundamental principles of plasticity theory, but
before attempting to apply these, chapter 6 is introduced. This is concerned
with the macroscopic behaviour of real engineering metals and is thus
intended to focus the attention on the complex behaviour of metals and
hence highlight the various limitations which are imposed when applying
any theory.
Chapters 7 to 9 inclusive endeavour to employ the preceding theory in
solving a number of metal forming problems which are likely to be of special
interest to the student and practising manufacturing engineer. These are
classified in terms of plane strain deformation, plastic strain with axial
symmetry and plane stress or pseudo-plane-stress problems.
Chapter 10 presents and discusses the so-called 'extremum' principles
for an ideal rigid perfectly plastic material leading to the formulation of
the lower and upper bound theorems. This latter theorem is particularly
useful in engineering applications because it can be employed to furnish
an acceptable overestimate of the important forming parameters.
I am conscious of the fact that the material presented in this text could
be adversely criticised. For example, limited reference is made to inertia
effects and the propagation of elastic and plastic waves upon the analyses
of the processes. Consequently, the treatment is essentially through a
quasi-static approach. No serious attempt is made to deal with the tribo-
logical aspects of metal forming, important as these are, and no reference
is made to the so-called high-energy forming methods or other forming
techniques such as hydro-forming. I apologise for these omissions but
emphasise the immense difficulties which are experienced when attempting
to cover all these factors in a single text. It is, however, hoped that readers
will be sufficiently stimulated by the subject matter of the present text to
give further consideration to such matters and they are advised to refer
to other literature specifically concerned with these topics. The references
and bibliography have, therefore, been made as extensive as possible.
My sincere thanks and appreciation are offered to Professor W. Johnson
formerly of the University of Manchester Institute of Science and Technology
and now at the University of Cambridge. As a friend and teacher, Professor
Johnson was responsible for introducing me to what I consider the most
interesting of studies and occupations. I am also indebted to him for his
stimulation by personal example in every respect.
The good influences of other past and present colleagues and students
alike are also acknowledged, and thanks are due to Mrs. Susan Mahoney
and Mrs. Ann Gray who typed parts of the manuscript.

London, 1977 R. A. c. s.
1
Introduction

1.1 DEFINITION AND SCOPE OF THE SUBJECT


When a solid body is subjected to a force of small magnitude it deforms
elastically such that the strain is directly proportional to the stress and
when relieved of the stress it eventually returns to its original dimensions.
Elastic deformation is therefore a reversible or recoverable process. The
well known theory of elasticity is concerned with the mathematical study
of stress and strain in elastically deformed solids.
Under the influence of a substantial force a solid body may experience
inelastic, plastic deformation which is an irreversible or irrecoverable process
and the body is permanently deformed. Actually, the so-called elastic body
is an idealisation because all solid bodies exhibit more or less plastic beha-
viour even when subjected to small forces. However, this permanent defor-
mation is so small as to be practically immeasurable.
The plastic properties of solids are extremely varied and depend on the
material under consideration, temperature, duration of the process and
other factors. The plastic deformation of ductile engineering metals such
as steel, aluminium, copper and brass at normal temperature is virtually
independent of time. However, metals at high temperatures experience
plastic deformation which increases with time, referred to as creep.
At the present time, the theory of plasticity is concerned with the mathe-
matical study of stress and strain in plastically deformed solids with particular
reference to metals. In recent years, the methods of plasticity theory have
been successfully projected in geophysical and geological problems. Materials
which are potentially capable of plastic deformation such as ice, clay and
rock have therefore been studied. It is usual when considering plasticity
theory to infer time-independent plastic deformation, that is, athermal
plasticity. Plastic flow in which time has an influence is studied in the theory
of creep, viscoplasticity and rheology.
The mathematical theory of plasticity is phenomenological in nature
and attempts to formalise experimental observations of the macroscopic
behaviour of a plastically deforming solid in a uniform state of complex
2 ENGINEERING PLASTICITY

stress. On the other hand, the physical explanation of the elastic and plastic
properties of solids, from a microscopic viewpoint, in relation to their
crystal structure is the subject of materials science. Nevertheless it is hoped
that the two approaches may eventually be merged into a unified theory of
plasticity.
Metal processing generally may be defined as that branch of engineering
which is concerned with the manufacture of components, assemblies,
machines and structures by the various processes of metal forming, machin-
ing, joining such as welding and bonding, and also casting of metal from
the liquid state. The content of this textbook is mainly confmed to analyses
of those metal forming processes in which the shape of the workpiece is
altered to the required shape by plastic deformation without material removal.
The general term 'workpiece' denotes the part on which the metal forming
process is performed. Specific terms such as 'billet' in extruding, 'blank'
in deep drawing are also used when discussing the specific processes. The
role of the metal forming specialist is to provide the designer and the manu-
facturing engineer with the information required to design and efficiently
operate metal forming equipment and control the products. The information
required includes the nature and extent of the deformation involved and
the forming parameters such as forming force, energy and power. It is thus
evident that for an understanding of the mechanics of metal forming proces-
ses, where plastic deformation is the principal mechanism by means of
which the workpiece is transformed into the desired shape, a knowledge
of the theory of plasticity is required.
The theory of plasticity can conveniently be divided into two ranges
depending upon the extent of plastic deformation. In one there are the
metal forming processes such as forging, drawing, extrusion and rolling,
etc., involving large plastic strains and where the elastic strains can be
considered as negligible. In the other there is a considerable number of
problems at the interface between elastic and plastic behaviour involving
small plastic strains of the same order as the elastic strains. However, these
elastoplastic problems, which are of considerable importance to the structural
designer, are not the subject of the present treatise.

1.2 NATURE OF ENGINEERING PLASTICITY


The main aims of the theory of engineering plasticity are twofold. The first is
to construct explicit relations between stress and strain for plastically deform-
ing solids such as metals and their alloys on the basis of experimental data.
The second is to develop mathematical techniques for determining non-
uniform stress and strain distributions so as to predict the plastic deformation
of a body in various circumstances. A particular feature of the theory is
that it is concerned with establishing laws of plastic deformation in a complex
stress state. These laws which agree satisfactorily with experimental evidence
have been established mainly for metals and their alloys. They may, however,
INTRODUCTION 3
retain their validity for other materials. Another characteristic of plasticity
theory is the nonlinearity ofthe principal laws and hence the basic equations.
The solution of some of these equations inevitably presents considerable
mathematical difficulties.
Engineering plasticity is concerned with the behaviour of metals during
plastic deformation and is applicable to a wide range of engineering problems.
These include the plastic bending of beams and plates, the overstraining
of spheres and cylinders which are widely used as pressure vessels in, for
example, the chemical industry and rotating discs, the optimal design of
structures and the metal forming processes. This treatise is devoted to the
application of engineering plasticity to some of the common metal forming
processes. In analysing the problems it is often extremely difficult to obtain
exact solutions because of the present mathematical intractability. As a
consequence, to obtain solutions which are industrially useful it is often
necessary to make certain assumptions and approximations and it may
not be possible to assess the resulting effects a priori.

1.3 A BRIEF IDSTORICAL ACCOUNT


It is generally regarded that the history of the science of the plasticity of
metals began in 1864 when Tresca 1 published a preliminary account of
his experimental results on punching and extrusion. His conclusions led
to the formulation of a yield criterion which states that a metal yields plasti-
cally when the maximum shear stress attains a critical value. One of the
earliest contributions to a mathematical theory of plasticity was made in
1870 by Saint-Venant 2 who applied the Tresca yield criterion to determine
the stresses in a partially plastic tube subjected to torsion or bending.
Saint-Venant also determined the stresses in a completely plastic tube
expanded by internal pressure and subsequently established a system of
five equations relating the stresses and strains in plane strain deformation,
that is, two-dimensional plastic strain. Following the ideas of Saint-Venant,
the three-dimensional relations between stress and plastic strain-rate were
proposed by Uvy 3 in 1870. Levy also introduced the method oflinearisation
for the plane strain problem.
Subsequent development of the theory of plasticity proceeded slowly
during the next 75 years. Some progress was made at the beginning of this
century when, in 1909, Haar and von Karman 4 obtained plasticity equations
from a variational principle. During the next decade, many experiments
were performed which were concerned with the yielding of tubes subjected
to various stress states. These experiments resulted in differing conclusions
and various yield criteria were suggested. For many metals, the most satis-
factory was a new yield criterion formulated by von Mises 5 in 1913 on the
basis of purely mathematical considerations. It is worth noting that a
similar yield criterion was given earlier although in a less precise form and
it was not actually related to the construction of the mathematical theory
4 ENGINEERING PLASTICITY

of plasticity. The von Mises yield criterion was interpreted later by Hencky6
as implying that yielding occurred when the elastic shear strain energy
attained a critical value. Stress-strain equations similar to those presented
by Levy were independently proposed by von Mises.
The plane plastic strain problem was shown by PrandtC in 1920 to be
hyperbolic and he determined the pressure required to indent a plane
surface by a smooth flat punch. Three years later, Hencky 8 produced the
general theory applicable to the special solutions by Prandtl and also defined
the geometrical properties of slip line fields for plane strain deformation.
Some time elapsed before Geiringer 9 obtained the velocity compatibility
equations for flow along slip lines. It is believed that the first application
of plasticity theory to technological processes was that by von Karman 10
who, in 1925, analysed the stress distribution during the rolling of metal
strip by an elementary method. Similar theories for wire drawing using
this so-called slab method were presented later by Sachs 11 and Siebel 12 .
It was not until 1926 that the Levy-von Mises stress-strain relations
were shown to be valid to a first approximation when Lode 13 carried out
experiments on various metal tubes subjected to combined tension and
internal pressure and measured the resulting deformation. However, his
results indicated certain discrepancies which were confirmed in 1931 by
the meticulous work of Taylor and Quinney 14• The theory of plasticity
was further generalised when Reuss 15 in 1930 introduced the elastic strain
component into the stress-strain relations and Schmidt 16 in 1932 and
Odquist 17 in 1933 indicated how the Levy-von Mises equations could
be modified to take into account the effect of strain-hardening.
The 1939-45 War stimulated research, especially in the USA and the
United Kingdom. An excellent account of the advances made in the period
up to about 1949 is given in the classical textbook on the mathematical
theory of plasticity by HilF 8 • The last twenty years have seen the establish-
ment and consolidation of a well defined knowledge in the plasticity of
metals with particular reference to both the design of structural elements
and the metal forming processes. Techniques for analysing and predicting
the forming parameters and deformation have been reasonably well defined
for a number of metal forming processes using the slip line field theory.
The theory is applicable to plane strain, plane stress and to axial symmetry.
It can be applied to isotropic materials, anisotropic materials and to materials
such as soil. However, it is the plane strain application which has been
most successfully exploited for metal forming processes. The number of
slip line field solutions has steadily increased during the last decade and
the problems solved are becoming more complex. The number of plane
stress and axisymmetric problems of practical importance that have been
solved is small.
A series of plane strain solutions presented in the early 1950s by A. P.
Green 19 - 2 \ for example, are notable achievements. Some other contri-
butions to plane strain slip line field solutions are those by Alexander 22
INTRODUCTION 5
and Johnson 23 .An extensive bibliography of published work concerned
with plane strain slip line field theory is given in the monograph by Johnson,
Sowerby and Haddow 24• The introduction of the hodograph or velocity
diagram by Prager25 in 1953 introduced a simplification into slip line field
solutions and a further innovation by Prager was the use of the cycloid
in the stress plane.
In 1951, Drucker, Greenberg and Prager 26 stated three so-called limit
theorems from which upper and lower bound estimates, that is, overestimates
and underestimates, may be established for the forming parameters in metal
forming processes. It is, however, recognised that these theorems were
deducible from the work principles published in the textbook by Hill 27 ,
and it appears that the earliest reference to the theorems of limit analysis
was probably due to Gvozdev in 1936. A translation of the paper from the
Russian is given by Haythomthwaite 28 • Numerous examples of the use
of the upper bound and lower bound theorems in the field of metal forming
processes are presented in the textbook by Johnson and Mellor 29 • Upper
bound solutions for plane strain problems were obtained by Kudo 30 by
introducing the concept of the unit rectangular deforming region. Several
types of admissible velocity field in the deforming regions were considered
and it appears that the rigid triangle velocity field is the best in which the
deformation zones are bounded by plane surfaces. The technique was later
extended by Kudo 31 to axisymmetric problems of forging and extrusion.
An improvement of the Kudo upper bound solutions for axisymmetric
problems was achieved by KobayashP 2 , who assumed the surfaces of
discontinuity to be curved instead of plane for the admissible velocity fields
in the unit deforming regions.

1.4 CLASSIFICATION OF METAL FORMING PROCESSES


Many different metal forming processes are utilised in the manufacturing
industry and there is difficulty in formulating a generally accepted classi-
fication of these processes. The following are some possible classifications:
(a) cold, warm or hot forming characterised by the homologous temperature
(see section 6.4.2); (b) chip forming or chipless forming characterised by
whether or not metal removal is involved; (c) state of stress in the workpiece,
that is, simple or complex; (d) type of stress involved, for example, tensile,
compressive or shear; (e) size of the plastically deforming zone which may
be local or general; (f) steady or non-steady state deformation; (g) low or
high strain-rate.
It will thus be appreciated that a metal forming process may not easily
be characterised by a single classification. Nearly all metal forming processes
involve the workpiece being subjected to complex stress states which can
vary from, say, triaxial compression to biaxial tension. However, shear
stresses need not be considered unless they constitute major stresses to
which the workpiece is subjected and are thus influential in contributing
6 ENGINEERING PLASTICITY

to plastic deformation. Most of the complex stress states can then be approxi-
mated by their principal stress components, that is, by the normal stresses
acting on planes on which shear stresses are absent.
Thomsen, Yang and KobayashP 3 have suggested, based on a scheme
originally proposed by Kienzle, that the kind of stress involved may be the
best choice, thus dividing the major industrial metal forming processes
into four main groups:

(1) Squeezing group in which the workpiece is subjected principally to a


compressive stress state. The processes in this group normally involve bulk
plastic deformation producing considerable change in the shape of the
workpiece. They include forging (upsetting, closed die forging and coining),
forward and backward extrusion, rolling, swaging, spin forging and rotary
forging.
(2) Drawing group in which the workpiece is subjected principally to a
tensile stress state; thus pulling instead of pushing is implied. The processes
in this group are generally limited in the extent of plastic deformation
of the workpiece which can be achieved in a single operation and are therefore
restricted to changes in shape of the workpiece rather than changes in
thickness. The workpiece is thus usually in the form of metal sheet, plate
or thin walled tubing. This group includes sheet, wire and bar drawing,
tube drawing, the deep drawing of cylindrical cup and box shapes and
stretch forming.
(3) Bending group in which the workpiece is subjected to couples thereby
inducing tensile stresses on one side of the workpiece and compressive
stresses on the other with a stress gradient throughout the thickness of the
workpiece. This group of processes are again restricted to change of shape
rather than change in thickness and include straight flanging, stretch flang-
ing (concave flanges), shrink flanging (convex flanges) and the seaming of
sheet or plate.
(4) Cutting group are those processes which either separate excess metal
from the workpiece in a single operation or by incremental metal removal.
The first category are shearing processes such as bar cropping, and piercing
and blanking of sheet metal which are regarded as chipless forming processes.
The second category are chip forming and include the conventional machining
processes such as turning, drilling, milling, grinding, sawing, broaching
and shaving.

In reference 33, illustrations of the metal forming processes are given


according to their classification within one of the four groups. The corres-
ponding approximate stress state in the plastically deforming zone and
the extent of the deformation zone are shown. The deformation is classified
as being steady or non-steady state for each of the processes.
The more recent methods of material removal including electrochemical
machining, electric discharge machining, ultrasonic machining, electron
INTRODUCTION 7
beam and laser machining can either constitute a fifth group or be included
in the cutting group of processes. It is emphasised that all the forming
processes which involve metal removal are considered beyond the scope of
this textbook.

1.5 FORMING LIMITS


The extent of plastic deformation that may be achieved during metal forming
processes is profoundly influenced by the stress state developed in the
workpiece. Plastic deformation can only occur if shear stresses are present.
Thus, a state of uniform triaxial tensile stress or uniform triaxial compressive
stress, referred to as a hydrostatic stress state, does not produce plastic
deformation irrespective of the magnitude of the stresses since shear stresses
are absent on any arbitrarily chosen plane. However, if the normal stresses
are unbalanced, that is, they are not all of the same magnitude, shear stresses
can exist and when a critical magnitude of shear stress is attained the initia-
tion of plastic flow occurs known as the yield condition. It is not important
whether yielding occurs due to the development of shear stresses resulting
from an unbalanced triaxial tensile stress state or an unbalanced triaxial
compressive stress state. Nevertheless, the extent of plastic deformation
that can be achieved is highly dependent on the nature of the stress state
induced.
Plastic deformation is limited by (a) necking which is a phenomenon due
to an instability condition in tension when uniform plastic flow ceases
and becomes localised resulting in local thinning of the workpiece, or
(b) buckling which is associated with a transition phenomenon between
elastic and plastic stress states, or (c) fracture which is a separation process,
or possibly by a combination of these defects. It follows that a limit to
forming is imposed when uniform plastic flow ceases and the forming limit
is determined by whichever defect occurs first. For example, in uniaxial
tension the fracture of ductile metals is usually preceded by local plastic
flow or necking. Consequently, the forming limit is determined by necking
rather than by fracture. A thin-walled tube subjected to torsion fails first
by buckling followed by fracture, whilst the torsion of a solid bar is limited
by fracture. The forming limits therefore depend on the state of stress induced
in the workpiece.
The condition of instability in uniaxial tension producing the onset of
necking occurs when the maximum axial force is attained and at a relatively
low strain. The deformation of the workpiece that can be achieved by the
processes in the drawing group is relatively low and is especially so if the
workpiece is subjected to a stretching operation as in stretch forming.
Since buckling is a transition phenomenon between the elastic and plastic
stress states, if it occurs it does so at an early stage in the forming process.
A typical example of this is the buckling or wrinkling of the flange at the
commencement of the deep drawing of a cylindrical cup. This limit to
8 ENGINEERING PLASTICITY

forming is usually suppressed by a change in the stress state induced in the


flange by means of the blankholder pressure.
The extent of the plastic deformation which can be achieved by those
processes comprising the squeezing group is relatively high. This can be
attributed to the stress state in the workpiece being predominantly triaxial
compressive which inhibits necking and fracture. In upsetting a cylinder
of ductile metal, such as copper, at ambient temperature a dimensional
change of 20: 1 can easily be obtained and even greater strains can be
induced locally when a metal billet is extruded.
In bending of sheet or plate, the outer fibres of the metal are stretched
whilst the inner fibres are compressed. Because of the stress and strain
gradients induced in the bent portion of the workpiece, the onset of necking
may be delayed or perhaps eliminated. However, fracture occurs in the
outer fibres if the radius of the bend is too small.
Considerable plastic deformation can be achieved for ductile metals
before separation by fracture occurs in the shearing processes where the
intention is to separate the required part from the workpiece by chipless
cutting. In the blanking process the sheared edge profile of the blanked
product is thus composed of a smooth portion due to plastic flow and an
irregular portion due to fracture. Fracture is initiated at the punch and
die comer profiles where high stress concentrations are developed.
In conventional machining, that is, the chip-forming, cutting processes,
extensive plastic deformation occurs at the junction ofthe undeformed work-
piece with that part which is transformed into the chip. Intense shear strain
occurs in a narrow region referred to as the shear zone. For ductile metals,
such as mild steel, no forming limit is reached and the chip is separated
from the workpiece in a continuous form. In the case of a less ductile metal,
such as cast iron, a discontinuous chip is formed which breaks into separate
pieces suggesting that the fracture limit is exceeded. It is interesting to
note at this point that continuous chip forming in high speed turning is
being discouraged for safety and economic reasons. Various methods of
chip breaking have been actively investigated in recent years.
If a hydrostatic pressure is superimposed on the normal stress state for a
forming process, necking can be eliminated and provided precautions
are taken to ensure that buckling does not occur the extent of plastic defor-
mation can be extremely high. By increasing the magnitude ofthe hydrostatic
pressure the initiation of fracture is suppressed. With the advent of high
pressure technology, higher hydrostatic pressures are being utilised in a
number of processes, for example, hydrostatic extrusion, which enables
otherwise difficult to form materials such as high speed tool steel and beryl-
lium to be successfully formed.
INTRODUCTION 9
REFERENCES
1. Tresca, H., Sur l'ecoulement des corps solides soumis a de fortes pression,
C. R. Acad. Sci. Paris, 59, 754 (1864)
2. de Saint-Venant, B., Memoire sur l'etablissement des equations
differentielles des mouvements interieurs operes dans les corps solides
ductiles au dela des limites ou l'elasticite pourrait les ramener a leur
premier etat, C. R. Acad. Sci. Paris, 70, 473 (1870)
3. Levy M ., Memoire sur les equations generales des mouvements interieurs
des corps solids ductiles au dela des limites ou l'elasticite pourrait les
ramener a leur premier etat, C. R. Acad. Sci. Paris, 70, 1323 (1870)
4. Haar, A. and von Karman, T., Zur Theorie der Spannungszustiinde in
plastischen und sandartigen Medien, Nachr. Ges. Wiss. Gottingen,
204 (1909)
5. von Mises, R., Mechanik der festen Korper im plastisch deformablen
Zustant, Nachr. Ges. Wiss. Gottingen, 582 (1913).
6. Hencky, H. Z., Zur Theorie Plastischer Deformationen und der hierduch
imMaterial hervorgerufenen Nebenspannunger, Z. angew. Math. Mech.,
4, 323 (1924)
7. Prandtl, L., Uber die Harte Plasticher Korper, Nachr. Ges. Wiss.
Gottingen, 74 (1920)
8. Hencky, H. Z., Uber einige statisch bestimmte Fiille des Gleichgewichts
inplastischen Korpern, Z. angew. Math. Mech.,3, 241 (1923)
9. Geiringer, H., Beitrag zum Vollstiindigen ebenen Plastizitiits problem,
Proc. 3rd Intern. Congr. App. Mech., 2, 185 (1930)
10. von Karman, T., Beitrag zur Theorie des Walzvorganges, Z. angew.
Math Mech., 5, 130 (1925)
11. Sachs, G., Beitrag zur Theorie des Ziehvorganges, Z. angew. Math.
Mech., 7, 235 (1927)
12. Siebel, E., The plastic forming of metals (translated by J. H. Hitchcock),
Steel, (Oct. 16, 1933 to May 7, 1934)
13. Lode, W., Versuch iiber den Einfluss der mittleren Hauptspannung auf
das Fliessen der Metalle Eisen, Kupfer und Nickel, Z. Phys., 36,
913 (1926)
14. Taylor, G. I. and Quinney, H., The plastic distortion of metals, Phil.
Trans. R. Soc., A230, 323 (1931)
15. Reuss; A., Beruecksichtigung der elastischen Formaenderungen in der
Plastizitiitstheorie, Z. angew. Math. Mech., 10, 266 (1930)
16. Schmidt, R., Uber den Zusammenhang von Spannungen und
Formanderingen im Verfestigungsgebiet, lng.-Arch. 3, 215 (1932)
17. Odquist, F. K. G., Die Verfestigung von flussiesenahalichen Korpern,
Z. angew. Math. Mech., 13, 360 (1933)
18. Hill, R., The Mathematical Theory of Plasticity, Oxford University
Press, London (1950)
19. Green, A. P., A theoretical investigation of the compression of a ductile
material between smooth flat dies, Phil. Mag., 42,900 (1951)
10 ENGINEERING PLASTICITY

20. Green, A. P., Plastic yielding of notched bars due to bending, Q. Jl


Mech. appl. Math., 6, 223 (1953)
21. Green, A. P., On symmetrical extrusion in plane strain, J. Mech. Phys.
Solids., 3, 189 (1954)
22. Alexander, J. M., The effect of Coulomb friction in the plane strain
compression of a plastic-rigid material, J. Mech. Phys. Solids, 3, 233
(1955)
23. Johnson, W., Extrusion through square dies oflarge reduction, J. Mech.
Phys. Solids, 4, 191 (1956)
24. Johnson, W., Sowerby, R. and Haddow, J. B., Plane Strain Slip Line
Fields: Theory and Bibliography, Edward Arnold, London (1970)
25. Prager, W., A geometrical discussion of the slip line field in plane
plastic flow, Trans. R. Inst. Techno/., Stockholm, 65, 27 (1953)
26. Drucker, D. C., Greenberg, W. and Prager, W., The safety factor of an
elastic plastic body in plane strain, Trans. Am. Soc. mech. Engrs, 73,
J. appl. Mech., 371 (1951)
27. Hill, R., The Mathematical Theory of Plasticity, ch. III, p. 60, Oxford
University Press, London (1950)
28. Gvozdev, A. A., The determination of the value of the collapse load for
statically indeterminate systems undergoing plastic deformation (trans-
lated by R. M. Haythornthwaite), Int. J. mech. Sci., 1, 322 (1960)
29. Johnson, W. and Mellor, P. B., Engineering Plasticity, ch. 13, p. 415,
Van Nostrand Reinhold, London (1973)
30. Kudo, H., Study on forging and extrusion. Part !-Analysis on plane
strain problems, Koku-Kenkyu-sho Shuho, Tokyo University, Tokyo,
Japan, 1 (1), 37 (1958)
31. Kudo, H., Some analytical and experimental studies of axisymmetric
cold forging and extrusion, Parts I and II, Int. J. mech. Sci., 2, 102;
3, 91 (1961)
32. Kobayashi, S., Upper bound solution of axisymmetric forming
problems- I, presented at the Prod. Engng Conf. of Amer. Soc. mech.
Engrs. (May 1963)
33. Thomsen, E. G., Yang, C. T. and Kobayashi, S., Mechanics of Plastic
Deformation in Metal Processing, ch. 1, Table 1.1, p. 4, Macmillan,
New York (1965)
2
Stress Analysis

2.1 SPECIFICATION OF STRESS AT A POINT

2.1.1 Internal forces


The internal forces present in a continuous body represent interaction
between molecules and ensure the existence of a solid body. These forces
act when no external forces are applied to the body but are not in themselves
the object of study here. However, under the action of external forces a
body will deform and the mutual position of molecules changes and also
the distances between them. The action of external forces that produce
deformation gives rise to additional internal forces. Because the study of
elastic and plastic deformation is primarily concerned with the external
forces involved it is necessary to convert the internal forces into external
forces. This is achieved by employing the so-called method of sections.

2.1.2 Contact and body forces


A state of stress is produced in a body when a system of forces acts upon it.
These forces can be applied in two ways: they can act on every particle
of the body, for example, gravitational and magnetic forces, or they can
act directly on an external surface of the body and be transmitted indirectly
to the interior of the body through its constituent particles. The first type
of force is referred to as a body force and the latter type may be referred
to as a surface force or contact force. The theories of elastic and plastic
deformation are mainly concerned with contact forces in static equilibrium
and these forces are considered to be uniformly distributed over areas of
the body but, for simplicity, are usually considered in terms of a single
resultant force.

2.1.3 Stress at a point in a continuous body


Consider a body to be subjected to a system of external forces in equilibrium,
12 ENGINEERING PLASTICITY

~">
~,
____/s
. ., I ~
I A ' , rPart removed
~ .............._ I/ cSF '
\ .
""i ' Section plane
I _ _,!.,-r-7-r-7'-tl.
I
I
I

F.,
Figure 2.1. Internal forces in a continuous body

F 1 , F 2 ... F 8 , as illustrated in figure 2.1. Now imagine that the body is


intersected by a plane so that the body is divided into two parts, A and B.
An intersecting curved surface may be considered instead of a plane. This
plane intersects the lines of action of the forces of interaction between
molecules situated adjacent to both sides of the plane. If one of the parts,
say part A, is removed, the system of forces applied to the molecules of the
part B located adjacent to the sectioning plane will not be balanced. The
remaining part, B, however, is in equilibrium. Hence, the system of internal
forces brought about by the method of sections and converted now into
external forces relative to part B must be balanced by the system of remaining
external forces F 5 ,F6 ,F7 and F 8 acting on this part of the body. Let the
resultant of the continuous distribution of internal forces exerted by part
A on part B in the sectioning plane be F. Similarly, the resultant of the
continuous system of internal forces exerted by part B on part A in the
sectioning plane will have the same magnitude F but has opposite sense.
An elemental area bA can now be isolated in the sectioning plane, that
is, very small compared with the dimensions of the section but still great
compared with the distances between individual molecules of the body.
This area is intersected by a large number of lines of action of the internal
forces applied to molecules of the part B and exerted by the removed part A.
Let bF denote their resultant force vector. The ratio bF jbA is known as
the average stress of internal forces in the body on the area bA.
For purposes of subsequent mathematical analysis the contour of the
elemental area is contracted around any of its points, say P, then the area
bA and the resultant force bF will diminish indefmitely. The limit of the
ratio
lim [bF jbA] = dF fdA= ur . (2.1)
cJA--+0
STRESS ANALYSIS 13
is called the stress at the point P in the r direction and in the plane of the
section considered. The important thing to note here is that the stress must
be referred to a particular plane. For any other plane passing through the
same point Pit will be appreciated by reference to figure 2.1 that the force
distribution on the plane and hence the stress will be different.

2.1.4 Stress components


Of course, the direction of the stress, a r, need not be normal to the reference
plane considered. It is therefore convenient to resolve the force ~F into two
components, one having a direction normal to the reference plane, referred
to as ~FN• and one having a direction tangential to the reference plane
which is the shear force ~F8 • The resolution of the force ~F into normal
and shear force components acting on an elemental area ~A is shown in
figure 2.2.
The normal stress, a, is then
lim [~FN/~A]
bA-o

which is considered positive when tensile and hence negative when com-
pressive in nature. The shear stress
lim [~F8 /~AJ
bA-o

is designated r.
However, to completely specify the stress at a point it is necessary to
specify the component stresses at that point acting on three orthogonal
planes passing through the point. The stress acting on any arbitrary plane
through the point can then be determined in terms of the stress components
acting on the three specified orthogonal planes.

Figure 2.2. Normal and shear force components of the resultant force bF acting
on an elemental area bA at a point P
14 ENGINEERING PLASTICITY

(a)

fry

X
z z
frz frz
(b) '~"zx (c l '~"zy

"Xz '~"yz

frx frx fry fry

'~"xz "Yz
X 0 0 y
'~"zx '~"zy
frz frx frz
'~"xy
0 y
'~"yx

(d) fry fry

'~"yx

'~"xy

frx
X
Figure 2.3. Convention for stress components referred to a cartesian system
of axes

For this purpose, it is often found convenient to use the cartesian system
of coordinate axes when the three mutually orthogonal planes are XOZ,
YOZ and XOY as shown in figure 2.3(a). The component stresses acting
on these planes at their point of intersection are then shown, in detail,
in figures 2.3(b), (c) and (d), respectively, and are all shown as being positive.
A double subscript notation is employed to define the direction of each
of the normal stress components a, and the shear stress components -r.
The first subscript designates the direction of the normal to the plane on
which the stress acts and the second subscript designates the sense of the
stress. Thus, axx• denotes a normal stress acting on the face of the element
which is parallel to the YOZ plane, the normal to which is parallel to the
x axis and the sense of the stress is in the OX direction. For simplicity,
it is customary to delete the second subscript for normal stresses so that
axx• ayy and azz become ax, ay and az, respectively. Also, 'txy• is a shear
stress component acting on the face of the element which is parallel to the
STRESS ANALYSIS 15
YOZ plane, the normal to which is parallel to the x axis and the sense of
this shear stress is in the OY direction.
It will now be seen that the complete specification of the stress at a point
is defined by the nine component stresses

Figure 2.4. Stress components referred to the cylindrical system of coordinates

Figure 2.5. Stress components referred to the system of spherical polar


coordinates
16 ENGINEERING PLASTICITY

Other coordinate systems may be employed to specify the stress compo-


nents at a point in a body. These include the cylindrical and spherical polar
coordinate systems. If instead of the cartesian coordinate system either
cylindrical or spherical polar coordinates had been used to define the stress
at a point in a body, then the equivalent component stresses would be as
illustrafed in figures 2.4 and 2.5, respectively.
The complete specification of the stress at a point in cylindrical coordinates
is then defmed by the nine component stresses

and in spherical polar coordinates by the nine component stresses

2.2 DIFFERENTIAL EQUATIONS OF EQUILIBRIUM IN


THE NEIGHBOURHOOD OF A POINT

2.2.1 Force equilibrium


Consider an infmitesimal element in the form of the parallelepiped with
its faces orientated parallel to the coordinate planes with reference to the
cartesian coordinate system as shown in figure 2.6. The element is isolated
in a body and surrounds a point P. Let the edges of this parallelepiped be
~x.~y and ~z. Its volume is then ~V = ~x.~y.~z and its mass is p~V =
p~x. ~y. ~z where p is the density of the body at the point. The components
of the stress in the vicinity of the point are then as shown in figure 2.6. In
addition to these stress components resulting from the external contact
forces exerted on the body there may exist body forces due to, for example,
gravitational or magnetic forces. Such body forces can be resolved into
the following three components in the OX, OY and OZ directions, res-
pectively:
Bx = Xp~x.~y.~z}
By= Yp~x.~y.~z (2.2)
Bz = Zp~x.~y.~z
where X, Y and Z are the component body forces per unit mass in the
three directions.
STRESS ANALYSIS 17
z
~'Tzy r
Tzy+-
~z
.oz
~'Tyz ..
'T.yz+-
l>y
• oy

~'7Xy cS
'Txy+ bx • x

Figure 2.6. Force and couple equilibrium in the neighbourhood of a point


in a body

Furthermore, the body may have an accelerated motion so that if the


component displacements are u, v and w in the OX, OY and OZ directions,
respectively, the component accelerating forces are
82 u
Ax= pbx.by.bz ot 2
82 v
AY = pbx.by.bz ot 2 (2.3)

82 w
Az = pbx.by.bz ot 2
Resolving forces exerted on the element in the OX direction gives

( (TX + ~:X bX )bybz- (TXbybz + ( 1:YX + O;;X by )bxbz- 1:YXbXbZ


+ ( rzx + o;;x bz )bxby- rzxbxby + Xpbxbybz = o( = pbxbybz ~:~) (2.4)
After simplifying and dividing by bxbybz
8ax 8ryx 8rzx
ax + oy + az +
X _
P-
o( _ 8otu)
- P
2
2 (2.5)

that is LFx = 0( = 02
p Ot2u bV )
18 ENGINEERING PLASTICITY

If the forces exerted on the element are also resolved in the OY and OZ
directions, respectively, then the following three differential equations
of force equilibrium are obtained:

or

(2.6)
or

or

In tensor notation, equations (2.6) can be expressed as

(2.7)

During plastic deformation there are many cases where body 1 forces
are insignificant or absent and where the motion can be regarded as 'steady
motion without inertia forces being involved. In such cases, equations (2.6)
reduce to

OO"x + OT:yx + OT:zx = 0


ax oy oz

OT:xy + GUY + OT:zy = 0 (2.8)


ax oy oz

In tensor notation, equations (2.8) can be expressed as


auijjoxi =0 (2.9)
If body forces and inertia forces are insignificant or absent then the
corresponding differential equations offorce equilibrium referred to cylindri-
cal coordinates can be shown to be
STRESS ANALYSIS 19
8ar 1 8-riJr 8-rzr (Jr- (JIJ 0
-a;:-+-;:- 8(} + 8z +-r-=

8-rriJ 1 8aiJ 8-rziJ 2-rriJ


-+--+-+
8r r 8(} 8z r
=0 (2.10)

8-rrz 1 8-riJz 8az Trz


-+--+-+
8r r 8(} 8z r
=0

and when referred to spherical polar coordinates

(2.11)

2.2.2 Couple equilibrium


For simplicity, assume that body and inertia forces can be disregarded
and consider the moments of forces exerted on the element about a line
passing through the central point P and parallel to the OX axis. If the
resultant couple exerted on the element about this line is zero, then

( Tzy + 8-rzy
8z uz
s: ) s: s: (jz
uXuy 2 + Tzyuxuy 2- ( Tyz + 8-ryz
s: s: (jz
8y uy
s: ) s: s:
uXuZ 2by
.:5y
- Tyz(jx(jz 2 =0 (2.12)

Neglecting quantities of the fourth order and simplifying gives Tzy - Tyz = 0,
that is, Tzy = Tyz.
By considering moments of the forces exerted on the element about a
line passing through the point P and parallel to the OY and OZ axes, res-
pectively, two other similar equations are obtained:

Tzy = Tyz)
Txz = Tzx
(2.13)
Tyx = Txy
or (Jij = (Jji
Equations (2.13) express the condition that, if a pair of orthogonal shear
stress components are present at the point P on one plane, there will be
complementary shear stress components on the other planes to establish
20 ENGINEERING PLASTICITY

equilibrium with regard to rotation. For the complete specification of the


stress at a point, the number of independent stress components to be defmed
is then reduced from nine to six, that is, three normal stresses and three
independent shear stresses.

2.3 THREE-DIMENSIONAL STRESS ANALYSIS

2.3.1 Resultant stress on an oblique plane inclined to


the three cartesian axes
Let the area of the oblique plane ABC shown in figure 2.7(a) be unity, then
area OBC = l
OAC=m
OAB=n
where l, m and n are the direction cosines of the normal to the oblique
plane relative to the axes OX, OY and OZ, respectively.
The resultant stress on the oblique plane ABC is denoted ~ which has
a direct stress component sn normal to the plane ABC and a shear stress
component s. tangential to the plane ABC. The components of sR are sx,
sY and sz in the direction, OX, OY and OZ, respectively. Resolving forces
exerted on the infinitesimal tetrahedron in the OX, OY and OZ directions,
respectively,

z
Area ABC=unity

(a) (b)

Figure 2.7. Stresses on an oblique plane


21

(2.14)

The resultant force exerted on the oblique plane is


sR = [s; + s; + s;] 112 (2.15)
which is illustrated in figure 2. 7(b). Since the area of the oblique plane
ABC is unity, sR is also the resultant stress on this plane.

2.3.2 Normal stress on the oblique plane


Resolving forces normal to the oblique plane produces
sn = sxl + sym + szn
Substituting for sx,sy and sz from equations (2.14) gives
sn = a) 2 + 7:xylm + rx)n + 7:yxml + aym 2 + 7:yzmn + 7:zxnl + 7:zynm + azn 2
(2.16)
Combining the three equal pairs of shear stress terms such as rxylm +
ryxml = 2rxylm, etc., then produces
sn = axP + aym 2 + azn 2 + 2(rxylm + ryzmn + rzxnl) (2.17)

2.3.3 Shear stress on the oblique plane


szR = szn + szs (2.18)
Substituting for sR from equation (2.15) and sn from equation (2.17) enables
the shear stress s5 on the oblique plane to be determined.

2.3.4 The cartesian stress tensor


In general, a set of nine quantities referred to one set of axes and transformed
to another set is a tensor of the second order. Thus, equation (2.17) is the
defming equation of the cartesian stress tensor and the matrix of the tensor
is given by its nine stress components

(2.19)

A particular stress component can be obtained by assigning the subscripts


x,y,z = i,j. All nine stress components of the stress tensor are obtained
by cyclic permutation of the subscripts, that is, letting i be the row of a
22 ENGINEERING PLASTICITY

3 x 3 matrix (direction of the normal to the plane on which the stress acts)
andj be the column (sense of stress action).
The stress tensor of equation (2.19) is a symmetric tensor because of the
equality of shear stresses. However, there are some peculiar conditions
for which the stress tensor will not be symmetric, as in the case when body
moments are present 1•

2.3.5 Principal stresses


If the directions of the vectors sR and sn coincide, then the shear stress,
s., on the oblique plane is zero. The oblique plane is then a principal plane,
sn is a principal stress and its direction is a principal direction. At every point
in a body there are at least three principal directions.
The stress Sa then has the same direction cosines l, m and n as the normal
stress component sn. The components of sn in the OX, OY and OZ directions
become

(2.20)

and if these values of sx, sY and sz are substituted into equations (2.14)
(O'x- sn)l + 'ryxm + 'tzxn = 0}
(O'y- sn)m + 'txyl + 'tzyn = 0 (2.21)
(O'z- sn)n + 'tyzm + -r:xzl = 0
To determine the principal stress represented by sn, it is necessary to
eliminate the direction cosines l,m and n from equations (2.21). Rewriting
equations (2.21) gives

(O'x- sn)l ~ 'tyxm + 'tzxn:


'txyl + (O'y sn)m + 'tzyn- 0
0} (2.22)
-rxzl + 'tyzm + (O'z- sn)n = 0
thus indicating that the equations are consistent in l, m and n. Equations
(2.22) can be represented in tensor notation by replacing l, m and n by
lx, ly and lz, then
(O'ij - c5ijsn)li = 0 (2.23)
For equations (2.22) to have a nontrivial solution for l, m and n, the determi-
nant of the coefficients must vanish, resulting in
O'x- sn)
(O'ij - (jijsn) = [ 'rxy (2.24)
'txz
STRESS ANALYSIS 23
If the determinant is expanded by using the Rule ofSarrus (see appendix 1):
(ax- sn)(ay- sn)(O"z- sn) + 't"yx't"zy't"xz + 't"zx't"xy't"yz- 't"xz(O"y- sn)'t"zx
- 't"yz't"zy(ax- sn)- (az- sn)-rxy't"yx = 0 (2.25)
Therefore
- S~ + S~(O"x + O"y + O"z)- sn(O"xO"y + O"yO"z + O"zO"x) + O"xO"Pz + 2-rxy't"yz't"zx
2
- 't"xzO"y 2
+ Sn't"zx- 2
't"yzO"x 2
+ Sn't"yz- 2
't"xyO"z 2
+ Sn't"xy =0
or s;- S~(O"X + (Ty + O"z) + Sn {(TX(TY + (Ty(Tz + (TZ(TX- (-r;y + -r;z + 't";X)}
- O"xO"yO"z- 2-rxy't"yz't"zx + (-r;zax + -r;xay + 't"~yaz) = 0 (2.26)
which can be written as
(2.27)
where
(2.28)

2.3.6 Stress invariants


If, instead of the axes OX, OY and OZ, a different set of axes OX', OY'
and OZ' had been chosen at the point 0, then the equation (2.27) for deter-
mining the principal stresses would remain the same except that the co-
efficients J 1 , J 2 and J 3 would be defined in terms of the stresses a~, a~ and
a~, etc., corresponding to the different set of axes.

Hence J1 = 0"~ + 0"~ + 0"~


J 2 = - ( 0"~0"~ + 0"~0"~ + 0"~0"~) + ...
and J 3 = 0"~0"~0"~ + ...
Equation (2.27) is a cubic equation in sn which has three real roots and
consequently there are at least three principal stresses which will be designa-
ted a 1 , a 2 and a 3 , respectively. Since the state of stress at a point is defined
uniquely by the principal stresses at the point and since the roots of equation
(2.27) are the principal stresses when the directions of the coordinate axes
are chosen as the principal directions, the coefficients of equation (2.27)
must be invariant. This means that whatever the choice of axes chosen in
the body at the point under consideration, the coefficients must be the same.
Thus, J 1 =ax+ aY + az =a~+ a~+ a~ and similarly for J 2 and J 3 • These
coefficients are therefore known as the first, second and third invariants
of the stress tensor.
If the directions of the coordinate axes coincide with the principal direc-
tions then there are no shear stresses on the principal planes and the stress
invariants reduce to
24 ENGINEERING PLASTICITY

J1=a1+a2+a3 }
J2 = - (a1a2 + a2a3 + a3a1) (2.29)
J3 = a1a2a3
When az = <yz = •zx = 0, the cubic equation (2.26) reduces to
s~- s~(ax + ay) + sn(axay- <;y) = 0
or s~ - sn(ax +a)+ (axay- .;y) = 0 (2.30)
which is the quadratic equation for the two-dimensional stress state, then

or (2.31)

By using the principal directions as reference axes the matrix of the stress
tensor becomes
0

~] (2.32)

Selection of the principal directions in a material subject to plastic defor-


mation frequently results in considerable simplification of the problem.
The subscripts 1, 2, 3 have been assigned arbitrarily to the three principal
stresses. Sometimes the subscripts are selected to indicate their relative
magnitudes.
Referring to the quadratic equation (2.31) for the two-ditnensional or
biaxial stress state, it can be seen that a 1 is the algebraically greater principal
stress and hence a 2 is the lesser. Hence, for a triaxial stress state a 1 , a 2 , a 3 :

2.3. 7 Some states of stress in terms of principal stresses


Certain states of stress may be visualised graphically by considering the
principal stresses acting on special kinds of infinitesimal volume elements.
This is illustrated in figure 2.8 where
(a) Triaxial state of stress, a1 =I= a2 =I= 0'3
(b) Cylindrical state of stress, a2=a3=f=a1
(c) Spherical state of stress, a1=a2=a3

2.3.8 Principal shear stresses and maximum shear stress


Assume that the coordinate axes are chosen to coincide with the principal
directions then the shear stresses referred to these axes are zero. The normal
STRESS ANALYSIS 25

Ia) (b) (c)

Figure 2.8. Particular states of stress in terms of principal stresses: (a) triaxial;
(b) cylindrical; (c) spherical

stress and shear stress on some oblique plane having direction cosines with
respect to these axes of l,m and n are given by equation (2.17) and equation
(2.18~ respectively.

Sn = u1[2 + u2m2 + u3n2 (2.33)


and s; = s~ + s~ + s~ - s~ (2.34)
However, when the shear stresses referred to the chosen coordinate axes
are zero, equations (2.14) show that
(2.35)
Substituting these values in equation (2.34) gives
s; = u~ 12 + u~m 2 + u~n 2 - (u 1l2 + u 2m 2 + u 3 n2 ) 2 (2.36)
It is already known that on the principal planes the shear stresses are a
minimum, that is, zero. The planes on which the shear stresses attain extre-
mum values are determined by obtaining the particular values of the direction
cosines l, m and n such that s. in equation (2.36) is a maximum. However,
there is a restriction on the values of the direction cosines, since [2 + m2 +
n2 = 1, and therefore only two of them can be independent.
Substituting for n 2 = 1 - 12 - m 2 into equation (2.36), differentiating
the resulting equation with respect to l and m and then equating these
derivatives to zero for a maximum or minimum, the following equations
are obtained:
l[ (u 1 - u 3 )l2 + (u 2 - u 3 )m 2 - !(u 1 - u 3 )] = 0 }
(2.37)
and m[(u 1 - u 3 W+ (u 2 - u 3 )m 2 - !(u 2 - u 3 )] =0
One obvious solution is given when l = m = 0 and then n = ± 1. When
m = 0, I =I= 0 then the first equation of (2.37) gives l = ± (1/2) 1' 2 and if l = 0,
m =1= 0 the second equation gives m = ± (1/2) 1' 2 • However, there are, in
26 ENGINEERING PLASTICITY

general, no solutions for 1= m i= 0 except in the special case when u 1 = u 2 •


By eliminating first 1and then m from equation (2.36), the following table
can be produced

0 0 ±1 0 ± (1/2)1/2 ± (1/2)1/2
m 0 ±1 0 ± (1/2)1/2 0 ± (1/2)1/2
n ±1 0 0 ± (1/2)1/2 ± (1/2)1/2 0

The first three columns give the direction cosines of the principal planes
on which the shear stresses are a minimum, that is, zero. The last three
columns give direction cosines for angles of ± n/4 such that these planes
bisect the angles between the principal planes. The shear stresses on these
planes will then have extremum values.
If these particular values of the direction cosines from the last three
columns of the table are substituted into equation (2.36) the extremum values
of the shear stresses are then given as

"C 1 : ± 1((J 2 - (J 3) }
•2- ± 2(u1- u3) (2.38)
•3= ±!(u1-u2)
which are usually referred to as the principal shear stresses. If u 1 is the
algebraic maximum principal stress, u 3 the algebraic minimum principal
stress and u 2 the intermediate principal stress such that u 1 > u 2 > u 3 , then
the maximum value of shear stress designated •max is given by
(2.39)
It will then be seen that the maximum shear stress, •max' acts on the
plane which bisects the angle between the planes of maximum and minimum
principal stress and is equal in magnitude to half the difference between
these principal stresses.
If the values of the direction cosines for the planes on which the principal
shear stresses occur are substituted into equation (2.33) the normal stresses
on these planes are given, respectively, as

N 1 = !(u2 + u 3) }
N 2 = !(ut + u3) (2.40)
N3=!(ut +u2)

2.3.9 Octahedral stresses


The use of the invariants of the stress tensor can offer advantages by reducing
the number of significant parameters necessary to represent the state of
STRESS ANALYSIS 27
stress in a plastically deforming body. The same purpose is served by the
introduction of the so-called octahedral stresses which, as will be shown,
have a close relationship with the invariants ofthe stress tensor.
The octahedral plane is defmed as that which is inclined at equal angles
to the three principal directions such that
l=m=n
and since 12 + m2 + n2 = 1
31 2 = 1
or l = m = n = 1/(3) 112
That is cos- 11=cos- 1m= cos- 1n = 54°44'.

Substituting these values in equation (2.36) gives

(2.41)

= [ a12 + ;22 + a3-


2 a12 + a22 + a32 + 2( ~1a2 + a2a3 + a3a1 >]1/2

= [2(a~ +;~+a~)_ 2(a1a2 + a~a3 + a3a1) J' 2 (2.42)

The shear stress on the octahedral plane which is referred to as the octahedral
shear stress, tocT• is thus given by
tocT = s"(l=m=n) = t[ (a 1 - a2)2 +(a2 -a3)2 + (a3 -a 1)2]112 (2.43)
Equation (2.41) can be written as

- [ (a 1 + a2 + a3) - 2(a1 a2 + a2a3 + a3a1)


2
t
OCT-
-s8(I=m=n)- 3
- (a1 +a~+ a3)2J/2

=[~(a 1 +a2 +a3)2 - j(a1a2 + a2a3 + a3a1>J 112


= [~J~ +jJ2]112
Hence, the octahedral shear stress expressed in terms of the invariants of
the stress tensor is
(2.44)

The octahedral normal stress is obtained by substituting l = m = n = 1/(3)1' 2


in equation (2.33) which then gives
28 ENGINEERING PLASTICITY

O"ocT = (atf3) + (a 2/3) + (a 3/3)


= (a1 + 11 2 + 11 3)/3
=(1m =Jtf3 (2.45)

2.3.10 Representative or equivalent stress


Another invariant parameter of the state of stress which is often used with
advantage is known as the representative or equivalent stress and is defined by

3
ii = (2)112 ·'ocT= [1{2 (a1 - a2) 2 + (a2 - a3) 2 + (a 3 - a 1f }]1/2(2.46)

= (~)1/2·(~Y~T J3i + 12 J/2


(2.47)
While the representative or equivalent stress cannot be visualised as
acting on a specified plane, as can the octahedral stresses, it has the advantage
that in the case of a uniaxial state of stress, a 1 =I= 0, a 2 = a 3 = 0, as occurs,
for example, in simple tension or compression, it becomes equal to the
non-vanishing principal stress a 1 .

2.3.11 Spherical and deviator stresses


In general, the deformation due to the stress state a 1 , a 2 , a 3 consists of
two components: (a) a volumetric or hydrostatic component, and (b) a
distortional component which produces a change in geometry of the body.
Since changes in volume are usually considered negligible during plastic
deformation, the components of stress responsible for change in shape are
of considerable importance.
Any stress state can be reduced to two components:
(1) the hydrostatic stress
(2.48)
where for a real material the mean stress am will produce the same volume
change as the three principal stresses a 1 , a 2 , a 3 ;
(2) the reduced or deviator stress
(2.49)
Hence
or a~= (2a1- a2- a3)/3 }
Similarly a~= (2a 2 - a 3 - a 1)/3 (2.50)
and a;= (2a 3 - a 1 - a 2)/3
STRESS ANALYSIS 29
2.3.12 Spherical and deviator stress tensors
Spherical and deviator stress tensors can be formed because the sum of
two tensors produces another tensor. Thus, in general,
(2.51)
where ! akk = !(11 1 + a 2 + a 3) = 11m and <\ is Kronecker delta.
Hence t<>ijakk may be defined as the spherical or hydrostatic stress tensor

!<\11kk = 0
~11m 0
11m
0J
0 =
ll110 0
11 2 = 11 1
0
0
J (2.52)
0 0 11m 0 0 11 3 =11 1
Then solving for a;j, the defining equation of the deviator stress tensor is

which in terms of the components of the stress tensor equation (2.19) is

n
given by


!xy 0
a;j = !yx
!zx
11y
!zy
'!yzuJ - r·m00
11z
11m
0
(2.53)

' - [<•. !yx- •,) !xy !xz


or 11ij-
!zx
{ay- am)
!zy
,,. J
(az- am)
(2.54)

It is now evident that the deviator stress tensor is composed only of


the deviator stress components, a~= (ax- am), 11~ = (ay- am) and 11~ =
(az -am) and the shear stress components •xy•'"'yz and •zx· It will be shown
later that plastic deformation can take place only when the deviator stresses,
shear stresses or both having a certain combined magnitude are present.

2.3.13 Invariants in terms of the deviator stresses


J 1 = 11 1 + 11 2 + 11 3 = {11~ + 11m) + {11~ + 11m) + {11~ + 11m)
= (11~ + 11~ + 11~) + 3am
= J'1 + J1
where J~ =(a~ +a~+ a~)
and (2.55)
lz = -(a1a2 + 11z113 + 113111)
= - {(a~ +am)( a~ +am)+ (a~ +am)( a~ +am)+ (a~ +am)( a~ +am)}
= - {11~ 11~ + am(a'1 + 11~) +a!+ 11~ 11~ + am(a~ + 11~) + a!
+ 11~11~ + am(a~ +a~)+ a!}
30 ENGINEERING PLASTICITY

However J~ = u~ + u~ + u; = 0
Therefore J 2 = - { (11I(12
I 11+11+32}
+ (12(13 (13(11 (Jm
=J~ -J~

where J~ =- (u~ u~ + u~u; + u;u~)


and J /12- - 3,.2 - 1.J2
vm- 3 1

and is a function of the hydrostatic stress only.


Therefore J2 = J~ - 1Ji or J~ = J 2 + 1Ji (2.56)
J 3 = u 1 u 2u 3 = (u~ + um)(u~ + um)(u; +urn)
= u~ u~ u; + um ( u~ u~ + u~ u; + u; u~)
+ u!(u~ + u~ + u;) + u!
Therefore J 3 = J;- umJ~ + u!J~ + u!
= J;- umJ~ + J~
where J; = U 1 u~ u;
1

u!J 1 = 01

and J /13 = (Jm3 = 271 J31 (2.57)


Then J3 = J;- fJ1(J2 + fJi) + 217J~
=J; -tJ1J2 -l7Ji
Therefore = 2\(2J~ + 9J1J2 + 27J3)
J; (2.58)
(u~ + u~ + u;) 2 = (J~) 2 = (u~) 2 + (u~) 2 + (u;) 2 + 2(u~ u~ + u~u; + u;u~)
but J~ = (u~ + u~ + u;) = 0 and- (u~ u~ + u~u; + u;u~) = J~
Therefore 0 = (u~) 2 + (u~) 2 + (u;) 2 - 2J~
J~ = H (u~f + (u~)2 + (u;)2} (2.59)
The third invariant can also be expressed as
(2.60)

2.3.14 Representation of a state of plane stress by the Mohr stress


circle diagram
A two-dimensional or plane stress state existing at a point in a body may
be represented graphically by means of the Mohr stress circle diagram 2
which can be constructed using a pencil, ruler, compasses and protractor
only. This approach can be particularly useful when, for example, rotation
of the system of axes is required. It is also valuable for illustrating certain
STRESS ANALYSIS 31

(b)

Figure 2.9. (a), (b) Representation of a state of plane stress at a point P: (a)
physical diagram for an arbitrary plane AC; (b) physical diagram for rotation of axes

facts and equations and can assist in the study of slip-line field theory which
is discussed in chapter 7.
For convenience, it will be assumed that the plane stress state existing
at a point P in a deforming body is present in the XOY plane and is as
shown in figure 2.9(a). Since the normal stresses present during metal forming
processes are predominantly compressive it will be assumed that the normal
stresses ux and uY are both compressive and that ux < uY algebraically.
An arbitrary plane AC is inclined at an angle 4>, anticlockwise, to the
plane BC on which the normal stress ux and the shear stress -rxy act. Let
u and -r be the normal stress and shear stress, respectively, acting on this
arbitrary plane with the senses as shown in figure 2.9(a).
Resolving forces exerted on the element ABC in a direction normal to
theplaneAC
uAC- uxBC cos 4> + 't"xyBC sin 4>- uYAB sin 4> + 't"yxAB cos 4> = 0
32 ENGINEERING PLASTICITY

or rJ- rJx cos 2 cp + Txy cos cp sin cp- rJY sin 2 cp + Tyx sin cp cos cp = 0
and therefore, rJ = rJx cos 2 cp + rJY sin 2 cp - Txy sin 2cp (2.61)
which can be rewritten as follows
rJ = !rJx(2 cos 2 c/J) + !rJY(2 sin 2 c/J)- Txy sin 2cp
= !rJx(cos2 c/J) + 1- sin 2 c/J) + !rJy(sin 2 cp + 1- cos 2 cp)- Txysin 2cp
= !(rJx + rJY) + !rJx(cos 2 cp- sin 2 c/J)- !rJy(cos 2cp- sin 2 c/J)- Txy sin 2cp
(2.62)
Resolving forces exerted on the element ABC in a direction parallel to
the plane AC
1:AC- TxyBC cos cp - rJxBC sin cp + TyxAB sin cp + rJYAB cos cp = 0
or 1:- Txy cos 2 cp - rJx cos cp sin cp + Tyx sin 2 cp + rJY sin cp cos cp = 0
and therefore, 1: = !(rJx- rJY) sin 2cp + Txy cos 2cp (2.63)
Squaring equation (2.62) produces
rJ 2 = {t(rJx + rJy} }2 + !(rJ; - rJ;) cos 2cp - (rJx + rJy)Txy sin 2cp
+ {t(rJx- rJy) COS 2cp }2 - !((Jx- (Jy)Txy Sin4cp + (Txy Sin 2cp) 2 (2.64)
Squaring equation (2.63) produces
1: 2 = {t(rJx- rJy} sin 2c/J }2 + !(rJx - rJy)Txy sin 4cp + (1:xy cos 2cpf (2.65)
Adding equations (2.64) and (2.65) then gives
(J 2 + 1: 2 = {t((Jx + (Jy) y + !((J;- (J;) COS 2cp- (rJx + (Jy)Txy Sin 2cp
+ {t((Jx- (Jy) y + 'l:;y
and therefore
rJ 2 - {t(rJx + rJy) Y- !(rJ;- rJ;) cos 2cp + (rJx + rJy)Txy sin 2cp + 1: 2
= {t((Jx- (Jy) Y + 'l:;y
Or rJ 2 - (rJx + (Jy){t((Jx + (Jy) + !((Jx- (Jy) COS 2cp- Txy Sin 2cp} + {t((Jx + (Jy) y + 1: 2
= {t((Jx- (Jy) y
+ 'l:;y
which reduces to
(J2 _ (rJx + (Jy)(J + {t((Jx + (Jy} }2+ 1:2 = {t((Jx _ (Jy) }2 + 1:;y
or {(J- !((Jx + (Jy) V + '1: 2 = {t((Jx- (Jy} Y + 1:;y (2.66)
and this can be written in the form
(rJ- Af + Tz = Rz (2.67)
which is the equation of a circle in the T-rJ plane having a radius
STRESS ANALYSIS 33
R = [ {!(ax- ay) }2 + .;Y] 1/2 (2.68)
with its centre at !(ax+ ay), 0.
The circle is the locus of points such as <,a in a stress plane for which the
coordinate axes are chosen to be the shear stress • as the ordinate and the
normal stress a as the abscissa.
To define the location of the points (<xy• ax) and (•yx' ay) on the Mohr
circle it is necessary to adopt a convention for the sense of shear stresses.
It will be assumed here that if the couple due to a shear stress is clockwise
in effect then the shear stress is considered positive. Thus, <yx is positive
whilst •xy is negative. Both ax and aY are compressive and therefore considered
negative and are plotted as abscissae to the left of the origin 0 in the Mohr
stress circle diagram of figure 2.9(c) such that OB represents the magnitude
of ax and OE represents the magnitude of aY to some scale. The shear stress
•xy• which is negative, is given by BD and <yx• which is positive, is given
by EF to the same scale. The resultant stresses on the planes PY and PX
are then represented by OD and OF, respectively. The radius of the circle,
R, is thus given by CD = CF.
The centre ofthe circle, C, is on the -a axis at !(ax+ ay) from the origin 0.
Therefore OC = f(ax + ay)

lei
Figure 2.9. (c) Mohr stress circle diagram
34 ENGINEERING PLASTICITY

Also CB =CD cos 2/3 = CP cos 2/3 = t(ax- ay)


and BD = 'xy =CD sin2f3 = CP sin2f3
From equation (2.62)
a= t(ax + ay) + t(ax- a) COS 24>- <xy sin 24>
= OC + CP cos 2/3 cos 24> - CP sin 2/3 sin 24>
= OC + CP(cos 2/3 cos 24> - sin 2/3 sin 24>)
= OC + CP cos 2(/3 + c/>)
=OC+CQ
or a=OQ (2.69)

From equation (2.63)

'=t(ax- ay) sin24> + <xy cos2cf>


= CP cos 2/3 sin 24> + CP sin 2/3 cos 24>
= CP sin 2(4> + /3)
or <=QP (2.70)
It should be noted that the corresponding angles in the Mohr stress
circle diagram of figure 2.9(c) are twice those in the physical planes of figure
2.9(a) and 2.9(b). Thus, the plane AC, where the stress state is (r, a), is inclined
in the physical plane at an angle 4>, anticlockwise, to the plane BC where
the stress state is (<xy• ax). This, however, is shown in the Mohr stress circle
diagram of figure 2.9(c) as 24> anticlockwise. The resultant stress on the
plane AC at the point Pis given as OP and is inclined at an angle t/1 to the
plane AC.
The principal stresses a 1 and a 2 where a 1 > a 2 algebraically are given
by the stress quadratic equation (2.31). However, for the case considered
here ax and aY are both compressive and consequently negative.

Therefore a
a 1 = _l.(a
2 X
+a)+
y -
[{.l(a
2 X
_ay )}2 + ,2xy ]1/2
2

= - f(ax + ay) ± R
hence a 1 = OC - CG = OG }
(2.71)
and a 2 = OC + CG = OC + CH = OH
in the Mohr stress circle diagram of figure 2.9(c).
These principal stresses act on orthogonal planes where the shear stresses
are zero. The shear stress, r, on any arbitrary plane AC is given by equation
(2.63). Thus, r = f(ax- ay) sin 24> + 'xy cos 24> = 0 if the plane AC is a
principal plane. Therefore, f(ax- ay) sin 24> = - 'xy cos 24>
STRESS ANALYSIS 35

or (2.72)

There will be two such planes given by

2cf> = tan -1{ - .!.( 'rxy


2
_
(]'X Uy
)}

Thus 2cf> = 1t - 2/J or cf> = ~- fJ }


(2.73)
and 2cf> = 2n - 2/J or cf> = 1t - fJ
measured anticlockwise from the PY plane on which the normal stress
ux acts.
Since 1: = f(ux- uy) sin 2cf> + -rxy cos 2cf>

:; = (ux- uy) cos 2cf> - 2-rxy sin 2cf>

= 0 if the shear stress is to have an extremum value.

Hence tan 2cf> = f(ux- uy) (2.74)


'rxy
Again, there will be two such orthogonal planes on which the shear stress
has extremum values given by

2cf> =tan -l{f(ux- ur)}


'rxy

Thus 2cf> = -1t - 2/J or cf> = -1t - fJ }


2 4
(2.75)
3n 3n
and 2cf> = - - 2/J or cf> = - - fJ
2 4
measured anticlockwise from the PY plane on which the normal stress
ux acts.
If the particular values of2c/> given in equations (2.75) are now substituted
in equation (2.63) it will be found that
'rmax = ± {t(ux- uy) cos 2/J + 'rxy sin 2/J}
However, sin2fJ = -rxrfR and cos2fJ = f(ux- ur)/R
Therefore -rmax = ± [ {t(ux- ur) }2 + -r;r]IR
= ± [ {t(ux- o) }2 + -r;Y]l/2
=±R
= ±f(ul -u2)
as illustrated in the Mohr stress circle diagram of figure 2.9(c).
36 ENGINEERING PLASTICITY

The maximum shear stress, 't'max• therefore acts on orthogonal planes


which bisect the angles between the planes of maximum and minimum
principal stress and is equal in magnitude to half the difference between
these principal stresses.

REFERENCES
1. Sokolnikoff, I. S., Mathematical Theory of Elasticity, 2nd ed., p. 42,
McGraw-Hill, New York (1956)
2. Mohr, 0., Abhandlungen aus dem Gebeit der technischen Mechanik,
2nd ed., p. 192, Wilhelm Ernst, Berlin (1914)
3
Strain Analysis

If the relative position of any two points in a continuous body is changed


then the body is said to be deformed or strained. When the distance between
every pair of points in a body remains constant during its movement then the
body is referred to as a rigid body. The displacements of a rigid body may be
either translations or rotations which are known as rigid body displacements.
The analysis of strain is the study of displacements of points in a body
relative to one another when the body is deformed and is, therefore, not
directly concerned with rigid body displacements. Since the analysis of
strain is essentially a geometrical problem it is unrelated to the material
properties. The specification of the strain at a point is consequently the
same for both elastic and plastic deformations.

3.1 INFINITESIMAL DEFORMATION


3.1.1 lnfmitesimal strain at a point
Consider two arbitrary points A and B in an unstrained body where B is very
close to A and after straining the points move to A' and B', respectively.
The distance AA' is the displacement of A and BB' is the displacement of B.
If the distance A'B' is exactly equal to AB then the displacement is one of
translation as might occur in the case of a rigid body. If, however, A'B'
does not remain equal to AB then there is a displacement of B relative to A
and a state of strain exists in the body.
In general, the strain will not be homogeneous but will be different at
different points. However, if a sufficiently small element of the body be
considered, such as the parallelepiped ACDBEFGH in figure 3.1(a), the
strain may be regarded as sensibly homogeneous. In this case it may be
assumed that parallel straight lines remain straight and parallel, plane
surfaces remain plane and that, for all straight lines having the same direction,
the ratio of alteration in length to original length will be the same. Two
parallel straight lines of the same length will then be equally elongated or
contracted.
38 ENGINEERING PLASTICITY

(X +6XI+\U-ul,
ly+6yl+lv+6v),
lz +6zl+lw+6w)
z L
lx+6x, y+<Sy, z+6z l
M

.Sz

A
(X,¥,Z)

Ia) (b)
Figure 3.l. Components of displacement at a point

With reference to a cartesian system of axes, let the coordinates of A in


the unstrained state be (x, y, z) and of A' after straining be (x + u, y + v, z + w).
Then, u, v and w are the projections of the displacement of A, that is, AA'
into the planes XOZ, XOY and YOZ parallel to the axes OX, OY and OZ,
respectively. It is assumed that these quantities are infinitesimal and that
they are continuous functions of the coordinates x, y and z.
When the point A moves to A' the side AC of the parallelepiped elongates
to A'C' as shown in figure 3.l(b) and has an angular movement JA'C'. The

Ia) (b)
z

Figure 3.2. Direct strains, engineering shear strain and rotation in the xz plane
STRAIN ANALYSIS 39
z

Figure 3.3. Direct strains and shear strains in the yz plane

side CD elongates to C'D' and has an angular movement KC'D' and the
side DB elongates to D'B' and has an angular movement LD'B'.
By referring to figures 3.2(a) and 3.4 it will be appreciated that the angle
JA'C' has a component parallel to the plane XOZ and also to the plane
XOY and similarly for all other angles. The displacement of B relative to A
is thus compounded of all these motions. For the sake of clarity, the displace-
ments shown in the various figures have been grossly exaggerated.
The coordinates of B are x + c5x, y + c5y and z + c5z and after straining
become (x + c5x) + (u + c5u), (y + c5y) + (v + c5v), (z + c5z) + (w + c5w), where
c5u, c5v and c5w are evidently the projections of the displacement of B relative
to A into the planes XOZ, XOY, YOZ parallel to the axes OX, OY and OZ,
respectively.
Since u is assumed to be a continuous function of x, y and z, (u + c5u) will
be the same function of (x + c5x), (y + c5 y), (z + c5z). Thus, if u = f(x, y, z),
(u + bu) = f {(x + t5x), (y +by), (z + bz)} and this latter expression is expan-
ded by employing Taylor's theorem:

(u + c5u) = f(x,y,z) + ~~ t5x + ~~c5y + ~~c5z +(terms in higher powers


of c5x,c5y and c5z)
Since u = f(x, y, z) is assumed to be a very small quantity, these latter
terms can be neglected.

Then au au au
c5u = -c5x + -c5y + -c5z
ax ay az
It is evident that (oufox) c5x is the component of c>u independent of c5y
40 ENGINEERING PLASTICITY

l>v
QYII~ Ox-..,-...._
u+~6x~r-~~-r~--~

Figure 3.4. Direct strains and shear strains in the xy plane

and oz and is the projection of C relative to A into the XOZ plane parallel
to the OX axis. Hence, oujox is the direct strain at A in the direction OX
which is denoted by exx. This nomenclature may be interpreted as the rate
of movement in the OX direction of a point on a line parallel to OX at A.
The part of bu depending on by alone, that is, (oujoy) by, is that part of the
displacement of B measured parallel to the OX axis resulting from the
angular movement of CD in the plane XOY, as shown in figure 3.4, and
oujo y is the rate of shear of planes parallel to OX and perpendicular to OY
and can be considered as the angular strain of CD denoted by exy· Similarly
(oujoz) bz is the displacement of B parallel to the OX axis resulting from
the angular movement of DB as shown in figure 3.2(a) and oujoz is the
angular strain of DB in the plane XOZ which may be denoted by exz.
The equations for the relative displacements bv and bw in the planes
XOY and YOZ parallel to the OY and OZ axes, respectively, are similar to
that for bu.
au au au
Then bu=- bx+-by+-bz
ax ay az
av av av
bv = - bx +-by+- bz (3.1)
ax ay az
OW OW OW
bw = - bx +- oy +- bz
ax oy oz
or in tensor notation
(3.2)
The tensor
STRAIN ANALYSIS 41
au au au
ax ay az
av av av (3.3)
eij =
ax ay az
aw aw aw
ax ay az
-

is called the relative displacement tensor and, as can be seen, is not generally
symmetric about its main diagonal, that is, it contains effects of a rigid
rotational motion about an axis passing through 0.
The change in the right angle F AC to the angle F' A'C' is shown in figure
3.2(a).
The line AC moves to A'C' and the line AF to A'F'. In moving from AC
to A'C' the line AC moves through an angle JA'C' = (J(zx' the projection
of which in the plane XOZ is aw;ax. Since the strain is considered to be
sensibly homogeneous, the line A'F' will be parallel to the line D'B' and
therefore the angular movement of AF will be equal to the angle MA'F' =
LD'B' = (J(xz, the projection of which into the plane XOZ is aujaz.
C'J
The angle (J(zx ~ tan (J(zx = A' J

aw bx
ax
au
bx +ax bx
aw
ax
1 au
+-aX
Therefore

F'M
Also angle (J(xz ~ tan (J(xz = A'M

au bz
az
aw
bz + 8z bz
au
az
1 aw
+-az-
42 ENGINEERING PLASTICITY

Therefore

By referring to figure 3.3 similar expressions can be deduced for the


shear strains eyz = ovfoz and ezy =ow joy which occur in the YOZ plane.
Similar expressions are obtained for the other shear strains occurring in
the XOY plane by referring to figure 3.4. These shear strains are exy = oufoy
and eyx = ov /ox.

3.1.2 Engineering shear strains


The difference in angles F AC and F' A'C' shown in figure 3.2(a) is therefore
given by (oufoz +owfox) which is referred to as the engineering shear strain
and designated c/Jzx. This expression is composed of two angular strains
exz = oufoz and ezx =ow/ox so that c/Jzx = exz + ezx. The significance of the
engineering shear strain c/Jzx is illustrated in figure 3.2(b).
Similarly, the engineering shear strain parallel to the XOY plane is
c/Jxy = (ovfox + oufoy) = eyx + exy and that parallel to the YOZ plane is
c/Jyz = (owfoy + ovfoz) = ezy + eyz•

Then

(3.4)

When considering angular movements it is necessary to adopt a convention


for deciding when the angular movement is positive or negative. The right-
hand screw rule is adopted here such that when viewing from the origin
along a given axis the direction in which a right-handed screw would rotate
is considered positive. For example, by referring to figure 3.2(a) it will be
seen that when viewing from the origin 0 along the OY axis, in the positive
direction, the projection of the line AF in the XOZ plane rotates clockwise
to A'F' and hence the angular strain exz is positive. It follows that the angular
strain ezx is negative.
Sometimes it is found convenient to consider half the engineering shear
strains, particularly when using tensor notation; then
1
Yxy = 2c/Jxy =
y OXov + au)
2\ 1
oy = 2(eyx + exy)
1
Yyz = 2,c/Jyz =
yaw
2\ oy + az = 2(ezy + eyz)
ov) 1 (3.5)

1
Yzx = 2,c/Jzx = 2\az + OX = 2(exz + ezx)
1(ou ow) 1
STRAIN ANALYSIS 43
The infinitesimal strain of the parallelepiped at the point A is thus defined
by three direct strains and three engineering shear strains and, for simplicity,
the direct strains are designated by using a single subscript:
au
ex= ax
Direct strains
av
ey=ay
OW
ez=az (3.6)
av au
c/Jxy =ox+ oy
OW OV
Engineering shear strains
c/Jyz =BY + OZ
au ow
c/Jzx =az+ ax
These equations were derived by Cauchy.
3.1.3 The rotations
It should be noted that it is impossible to produce a system of equations
inverse to equations (3.6), that is, to express nine components of equations
(3.3) in terms of six components of strain. The equations (3.6) are therefore
inadequate since the geometrical representation of the deformations at
a given point are incomplete.
Let the element considered at the point A be a cube (Jx = Jy = Jz) and
the direct strains ex= ey = ez = 0. Referring to figure 3.2(a), the line AE,
where CAE is 45°, is rotated through an angle, say, + wY that is, clockwise
about the OY axis when viewed from the origin 0 along the.QY axis in the
positive direction to take up its strained position A'E'.
ow ow
!: _ ) _ E'P _ Jz + ox Jx + Tz Jz
Then tan ( 4 wy - A'P - ou ou
Jx + ox Jx + oz Jz
OW ow
1 +-+-
ax oz
1 au au
+OX +a;
ow
1
+ax-
1 au
+a;
44 ENGINEERING PLASTICITY

when oujox = ovjoy = owjoz = 0 and bx =by= bz.


Therefore tan ( ~ - wY) ~ ( 1 + ~=) (1 - ~:)
ow ou
~1+--­
ox oz
if the product of partial derivatives is neglected.
1t
n ) tan 4 -tanwY
Also, tan ( - - w = ------
4 y 1t
1 + tan 4 tan wr

_1- tan wr
-1 +tan wr
,_, _1-w_y

'"'"'1 +wy

or 1- w
y ox ou)
~ (1 + w)y (1 +ow- oz
~ 1 + ow - ou + w ( 1+ ow - ou)
ox oz y ox oz
and w
y
(2 + owox - ou) ~ ou -
oz oz ox
OW

ou ow
oz ox
---

w ------,----,,-
OW ou
or ~
y
2+---
0X OZ

Therefore WY ¥-
~ ~: ~=) = ~ ( exz - ezx)
Similar expressions can be obtained for the rotations about the OX and OZ
axes by referring to figures 3.3 and 3.4, respectively.
However, it should be noted that rotations wx and wz are positive, if
clockwise, when viewed alofig the OX and OZ axes in the positive direction
for each case.
Hence

(3.7)
STRAIN ANALYSIS 45

« _au
xz-u

z z

1xz

Deformation Sh«ar strain Rigid body rotation

Figure 3.5. Showing that the general case of deformation consists of a shear
strain and a rigid body rotation

:u--: I
I
I
I
I
I
I
I
+

Simpl« shear Pur« shear Rigid body rotation

Figure 3.6. Distinction between simple and pure shear

The deformation for a two-dimensional case can be considered diagram-


matically to consist of a shear strain and a rigid body rotation as shown in
figure 3.5. The distinction between simple shear and pure shear is illustrated
in figure 3.6.
Every second order tensor can be resolved into a symmetric tensor and a
skew-symmetric tensor. It follows, therefore, that if the tensor eii is resolved
into symmetric and skew-symmetric parts, the symmetric part will represent
pure deformation whilst the skew-symmetric part will represent rigid body
rotations without deformation.
Let
or (3.8)
46 ENGINEERING PLASTICITY

Thus
ou ou ou ou
ox oy oz ox +~)~eu
Kou
ox ox 2 oz +ow)
ox
ov ov ov ov Kov
eij =
ox oy oz =ll;j= Kou +~)
oy ax oy oz +ow)
oy
ow ow OW Kou + ow)K~+ ow)
ow
OX oy oz oz ox 2 oz oy oz
0 Kou ox Kou
oy -~) _ow)
2 oz ox

K~- KOV- OW)


+w;j=
ox ou)
oy 0
oz oy
Kow _ ou)Kow _ ov) 0
ox oz 2 oy oz
[~
= }cPyx = Yyx
l.cfJ - yzx
}cPxy = Yxy
eY
}cPzy = Yzy
}c/Jxz = Yxz
}cPyz = Yyz
ez
]
l
2 zx-

+[
0 -wz roY
w. 0 -(!)X

-roY (!)X 0
(3.9)
where e1j is called the irrotational or pure strain tensor and w1j is the rotation
tensor. For pure deformation, equation (3.3) becomes
(3.10)
3.1.4 Cubical dilatation
Consider an infinitesimal parallelepiped of initial volume t5xt5yt5z
as in
figure 3.1 to be in a state of strain.
If shear deformations alone are produced without elongation or contrac-
tion of its edges, the unit change in volume of the cube will be a small quantity
of higher order compared with the shear deformations. Consequently, it
will be assumed, neglecting small quantities of higher order, that the change
in volume t5V will depend only on the elongations or contractions of the
edges t5x, t5 y
and t5 z.
The volume of the element after deformation is then given by
t5x(1 + ex)t5y(1 + ey)t5z(1 + ez)
= t5xt5yt5z(1 + ex)(1 + ey)(1 + ez)
= t5xt5yt5z(1 +ex+ ey + ez + exey + exez + eyez + exeyez)
STRAIN ANALYSIS 47
Neglecting the last four terms in parentheses, which are small quantities
of the second and third order, produces
bV = bxbybz(ex + ey + ez)
The dilatational strain, L1, is then defmed as the change in volume per
unit initial volume, thus

(3.11)

3.1.5 Inrmitesimal strain of a line element


In figure 3.1let the diagonal length AB of the parallelepiped before deforma-
tion = r, and after deformation A'B' = r + br, then
r2 = bx 2 + by2 + bz2
(r + br) 2 = (bx + bu) 2 +(by+ bv) 2 + (bz + bw) 2
= bx 2 + by2 + bz2 + ou 2 + bv 2 + bw2 + 2(bxbu + bybv + bzbw)
Hence (r + br) 2 - r2 = bu 2 + bv2 + bw 2 + 2(oxbu + bybv + bzbw)
or rbr = bxbu + bybv + bzbw if small quantities of higher order are
neglected.
Substituting for bu, bv and bw from equations 3.1 gives

rbr = bx[ :: bx + :~ by + :: bz J
+ b{:: bx + :; by+ :~ bz J
+ b{ ~= bx + ~y + ~; bz J
ou
=bx 2 -+by2 ov ow
-+bz2 -+bxby ov au)
( -+-
ox oy az ax oy

+ bxbz(:: + ~=) + bybz( ~; + :~)

If l, m and n are the direction cosines of AB, then bx = rl, by = rm and


bz=rn,

Therefore ou
br=r[ l2 -+m ov
2 -+n ow+ lm(ov
2 - -+-ou)
ox oy az ax oy

+In-+-
oz ax
-+-
( au ow) +mn(ow ov)J
oy az

The strain in the direction of AB is then given as


48 ENGINEERING PLASTICITY

(3.12)

Since equation (3.12) must be true for all values of l,m and n a necessary
condition that equation (3.1) represents a rigid body motion is that er = 0
where l,m and n =I= 0.
Therefore
ou- ov- ow -0
ox- iJy ----a;-
and ov + ou)=o
( ox oy
(~: + ~:)=o
( ow+ ov)=o
oy oz
OV ou
That is ox=- oy
ou OW
oz =-ox
OW ov
oy = - oz
or (3.13)
Thus for a rigid body motion the tensor e;j of equation (3.3) is skew-
symmetric.

3.1.6 Strain compatibility equations


The displacement of a given point in a deforming body is determined by
the three components u, v and w as continuous functions of x, y and z and
the deformation at the point is defmed by the six components ex, eY, ez,
<Pxy• </Jyz and </Jzx of equations (3.6).
If the three displacement components are specified, all six strain compo-
nents can be determined uniquely being expressed in terms of the first
derivatives of the displacement components. It may, however, be anticipated
that the six strain components cannot be defined arbitrarily and that certain
interrelationships must exist.
A problem is encountered in calculating the displacements from the
strains. Equations (3.6) provide six equations for the three unknowns u,
v and w. It is evident therefore that these equations will not have a unique
solution for arbitrarily chosen strains and that some restriction must be
imposed in order that equations (3.6) have a solution. The interrelationships
STRAIN ANALYSIS 49
which must exist between the strains in order that the body remains conti-
nuous after straining, that is, the displacements be continuous functions of
the coordinates are established by the strain compatibility equations which
can be classified into two groups. From equations (3.4)

(3.14)

and from equations (3. 7)


ox _ ou)
2wz = (ov oy (3.15)

Subtracting equation (3.15) from equation (3.14)


ou
2oy = cPxy- 2wz (3.16)

and adding equations (3.14) and (3.15) produces


ov
2ox = cPxy + 2wz (3.17)

OU OV . r 11
.
an d smce ex= ox'
ey = oy h
it 10 ows t at

:X ( cPxy- 2wz) = 2::~Y = 2~~ (3.18)

(3.19)

Hence (3.20)

and (3.21)

Also from equation (3.15)

iwz = ~(ov- ou)


oz oz ox oy
o( ov ow) o( ow ou)
= ox oz + ay - oy ox + oz

because the second and third terms on the right-hand side of this equation
cancel. Therefore
2owz- o¢yz o¢zx
oz -Tx-Ty (3.22)

From equations (3.20) and (3.21)


50 ENGINEERING PLASTICITY

~(o</>xy- iex) = 202Wz


oy ox oy oxoy
=~(2owz)
ox oy
= ~(2oey- o<l>xr)
ox ox oy
o2ex + o2ey = o2</>xy (3.23)
or
oy2 ox 2 oxoy
This is one of the relations of the first group. Two more equations of the
same form can be deduced.
From equations (3.21) and (3.22)

~(- o</>xy + 2 ~) = 2 02Wz


oz oy ox oyoz
=~(2owz)
oy oz
= ~( o</>yz - o<l>zx)
oy ox oy
or 2 o2ey = ~(o</>yz + o</>xy- o<l>zx) (3.24)
oxoz oy ox oz oy
This equation is one of the relations of the second group and two more
equations of this form can also be deduced. The following system of six
strain compatibility equations, which were first derived by Saint-Venant,
can now be defined:

(3.25)
STRAIN ANALYSIS 51
3.2 FINITE DEFORMATION

3.2.1 Finite displacement of a point in a continuous body


The displacements and their derivatives have previously been assumed
extremely small so that infmitesimal deformation was considered. If the
deformation is fmite, and the strains become larger, it is obvious that the
relationships previously developed for infmitesimal strains become increas-
ingly inaccurate. In this case, the strains will no longer be linearly related to
the derivatives of the displacement. Furthermore, the equilibrium equations
must be satisfied in the deformed body and should therefore be considered
in terms of the deformed coordinates if these are considerably different
from the undeformed coordinates.
There are two methods of describing the deformation of a continuous
body when the deformations are finite, namely, the Lagrangian and the
Eulerian. The former method uses the initial coordinates of each particle to
describe the deformation whilst the latter metnod uses the coordinates of
the particles in the deformed state to describe the deformation.
More exact relations corresponding to finite deformation will now be
developed to replace the strain components in equations (3.6) which charac-
terise infmitesimal deformation.
The projections of the deformed line element A'B' shown in figure 3.1(b)
on the coordinate axes are obtained by subtracting the coordinates of A'
from B'. Let the projections of A'B' on the OX, OY and OZ axes be c;, '1 and
1/J, respectively, then
c; = x + bx + u + bu - (x + u) = bx + bu}
Similarly '7 = by + bv (3.26)
and 1/J=Oz+Ow
Equations (3.1) can then be rewritten as
au) bx +-by+-bz
c; = ( 1 +- au ou
ax ay az

'1 =av ov) by+-


- bx + ( 1 +- ov c)z
ax oy az (3.27)

1/1 =ow ow ( ow)


- bx +-by+ 1 +- bz
ax ay az

3.2.2 Finite strain coefficients


Dividing equations (3.27) by the deformed length of the line element A'B' =
r + br gives,
52 ENGINEERING PLASTICITY

e
r + f>r

'1
--=m ov
= [ -l+ ( ov J-r-
1+-ovoy ) m+-n (3.28)
r + f>r 1 ox oz r + f>r

-"' -= n =[owl+
r + f>r ox 1
ow m+(1+ ow)n]-r-
oy oz r + f>r

where
+
r f>r 1 ' r f>r 1 ' +
_e_=l · -"'-=m · -"'-=n
r f>r 1 +
are the direction cosines after deformation and
f>x {)y {)z
l=-· m=-· n=-
r ' r ' r

(3.28)
were the direction cosines ofthe line element before deformation. Equations
can then be written as follows:

l =A-r-
1 r + f>r
r
m =B--
1 r + f>r
r
and n =C--
1 r + f>r

deformation process and can therefore be eliminated from equations


by squaring all three equations and adding
(3.28)
The direction cosines l1 , m1 and n1 are not known in advance of the

12
1
+m2 +n2 = 1= (A2 +B2 +C2)(-r-)2
1 1 r + f>r

h.h .
w tc gtves (r +rf>r)2 -- A2 + B2 + C2.
It can then be shown that

(r~ f>rr =A2 +B2 + C2

,2[ 1+2 :: +(::r +(!:Y +( ~:rJ +


m2[ 1+2 ;; +(:;r +(;;y +(~;YJ+
STRAIN ANALYSIS 53

aw + (au)
n [ 1 + 2 fu oz + (aw)z]
oz + (av)
2 2
2
fu +
2zm[(au +au au)+ (av + ov ov) + (ow ow)]+
ay ax ay ax ax ay ax ay
21n[(au + ou au)+ (av av) + (ow+ ow ow)]+
oz oz OX oz ox ox oz OX
2mn[(au au)+ (av + ov ov) +(ow+ ow aw)J
ay az az oy az oy ay az
-r-
The term ( r+~r) can then be wntten
2 .
as equal to (G + 1), say, because
in the first three terms of the right-hand side of this equation it will be seen
that there is the sum z2 + m 2 + n 2 = 1.

Therefore c~~ry -1=G


[ { (r + ~;) - r} + J-
1 1= G

(e.+ 1f -1 = G
e; + 2e. = G
where e. is the conventional or engineering strain of the line element AB.
Then (3.29)

where
ex=:~+~[ (:~y + (!:Y + (::rJ
eY :~ + ~ [ ( :~ + ( :~ y y y
+ ( ~; J

r r rJ
=

ez = ~; + ~[ ( :~ + ( :~ + ( ~;
(3.30)
ov au au au OV ov ow ow
=-+-+--+--+--
OX oy ox oy OX oy ox oy
8
xy

ow ov au au av ov OW ow
=-+-+--+--+--
8
yz oy oz oy oz oy oz oy oz
au OW au au OV OV OW ow
8 zx = oz + OX + OZ OX + OZ OX + fu OX
3.2.3 Finite strain tensor
Comparing equation (3.29) with equation (3.12) it can be seen that the
54 ENGINEERING PLASTICITY

coefficients of equations (3.30) are related to the components of the tensor

l~
1

eij = :eyx
2exy
eY
!~]
!eyz (3.31)
1
2ezx 2ezy ez
which is known as the finite or small strain tensor to distinguish it from the
infinitesimal pure strain tensor of equation (3.9).
If its components are known the strain at a given point in a body can be
determined in any direction defmed by the direction cosines l, m, n. Equation
(3.29) can then be denoted by
e; + 2er = 2f(l,m, n)
Therefore er2 + 2er - 2:j = 0
or er = - 1 ± (1 + 2!)112
= (1 + 2!)112 - 1
if the negative root is not considered.
Assuming l = 1,m = n = 0,
then
which only reduces to ex = ex if ex ~ 1.
Thus, if the derivatives are small so that their products can be neglected,
the expressions of equations (3.30) reduce to those previously obtained for
the infmitesimal strains of equations (3.6). However, the physical interpreta-
tion given to infinitesimal strains is no longer applicable.
Nevertheless, it is sometimes possible to treat problems involving large
strains using the equations for infinitesimal strains. This is possible if the
problem is considered incrementally, that is, a small step at a time, and after
each step the coordinates are changed to correspond to the deformed body.
In essence, the problem is then solved as a series of successive small strain
problems.

3.2.4 Principal strains


In the study of the stress at a point it was found that three mutually orthogonal
planes exist on which there are no shear stresses, that is, the principal planes.
In a similar manner, planes exist on which there are no shear strains. The
normals to these planes will not change orientation when the body is
deformed. Thus, a line element originally normal to such a plane will either
elongate or contract but will not change direction. The normal directions
to these planes are the principal directions and the corresponding strains are
known as principal strains.
Consider a line element r normal to the oblique plane ABC shown in
STRAIN ANALYSIS 55
z

Figure 3.7. Principal strain vector

figure 3.7. Upon straining, it is assumed that the line element changes length
by an amount br but its direction remains the same if ABC is a principal plane.
The components of r and br in the directions OX, OY and OZ are then
proportional, that is,

8 = br
r
= (e x = bu)
bx
= (e Y = bv)
by
= (ez = bz bw)
and then bu = ebx; bv = eby; bw = ebz (3.32)
Therefore, equations (3.2) become
bU = exbX + exyby ·r exzbz}
bv: eyxbx + eyby + eyzbz (3.33)
bW - ezxbX + Bzyby + ezbZ
Substituting for bu, bv and bw from equations (3.32) into equations (3.33)
gives
(ex- e)bx + exyby + exzbz = 0
eyxbx + (ey- e)by + eyzbz = 0 (3.34)
ezxbx + ezyby + (ez- e)bz = 0
that is ebxi = eijbxj
or (eij- bije)bxj = 0 (3.35)
The three equations (3.34) will have a non-vanishing solution only if the
determinant of the coefficients vanish.
Therefore

(3.36)
56 ENGINEERING PLASTICITY

If this determinant is expanded, a cubic equation is obtained which is


similar to equation (2.26) for stresses but with the stresses replaced by strains
as follows:
t: 3 - t: 2(t:x +By+ t:z) + t:{ t:xt:y + 8x8z + 8y8z- (a;y + B~z + a;x)}
- 8x8y8z - 2t:xy8yz8zx + (8x8~z + 8Yt:;x + t:zt:;y} = 0 (3.37)
or t: 3 - I 1t: 2 - I 2 t: - I 3 = 0 (3.38)
where the strain invariants are
I 1 = ex + By + Bz
I 2 = - (t:xt:y + 8x8z + t:yt:z) + (a;y + B~z + a;x) (3.39)
I 3 = 8x8lz + 2t:xy8yz8zx- (cx8~z + 8y8;x + 8z8;y)
If the three roots of the cubic equation (3.37) are the principal strains and
the oblique plane ABC of figure 3.7 is a principal plane, t:xy = Byz = t:zx = 0
and (a- t: 1)(t:- t: 2 )(t:- t: 3 ) = 0, where t: 1 , t: 2 , t: 3 are the principal strains.
The strain invariants in terms of the principal strains become
+ 82 + 83
I 1 = 81
I2 = - (8182 + 8283 + 8381) (3.40)

3.2.5 Principal shear strains and maximum shear strain


A direction exists at every point in a strained body for which the shear strain
is a maximum just as there is a direction for which the shear stress is a
maximum.
Let the system of cartesian axes be chosen to correspond to the principal
strain directions 01,02 and 03, respectively, and consider a line element
OP of unit length, as shown in figure 3.8, to have direction cosines l, m and n
with respect to these axes.

Figure 3.8. The geometry of strain


STRAIN ANALYSIS 57
Assume that the line element OP be subjected to a small strain so that
P moves to Q and the principal strains are 13 1,13 2 and 13 3 . It will then be seen
from figure 3.8 that the strain is composed of two parts: a linear or direct
strain 13 = RQ/OR and a shear strain() R: RP/OR.
The projections of OP on the 01, 02 and 03 axes are l, m and n, respectively.
After straining, these projections become /(1 + 13 1), m(1 + 13 2) and n(1 + 13 3)
where 13p 13 2 ,13 3 are the increases in length per unit original length in the
principal directions. The strained unit length is then given by
OQ2 = f(1 + 131)2 + m2(1 + 132)2 + n2(1 + 133)2
that is, (OR+ RQ )2 R: (OR 2 + 20R · RQ)
R: f(1 + 213 1) + m2(1 + 213 2) + n 2(1 + 213 3)
if the squares of small principal strains are neglected.
R: [2 + m2 + n2 + 2(131/2 +82m2+ 133n2)
R: 1 + 2(13 1 /2 +8 2 m 2 + 13 3n 2)
Dividing by (OR) 2 gives
1 + 2RQ/OR R: 1 + 2(13 1f +8 2m 2 + 13 3n 2)
Therefore 13 = RQ/OR R: 13 1/2 +8 2 m 2 + 13 3n 2 (3.41)

2 ()2 = (RQ)2 + (RP)2 = (PQ)2 R;; (PQ)2 "f () . 11


13 + (OR)2 (ORf 1 IS sma

Therefore 132 + ()2 R: 13i [2 + 8~m2 + 133n2 (3.42)


and () 2 ""8 2 2 2 2 2 2 22
~1 f+8 2 m +8 3 n -(8 1 f+l3 2 m +8 3 n ) (3.43)
In addition to equation (3.43) there is a restriction on the direction cosines
so that [2 + m 2 + n 2 = 1 and then only two of them can be independent.
Substituting n 2 = 1 - [2 - m 2 into equation (3.43) gives
()2 R: 13i [2 + 8~m2 + 13~(1 -[2-m2)- {131[2 +82m2+ 133(1 - [2- m2)}2
If this equation is now differentiated with respect to l and m, respectively,
and the derivatives equated to zero, the direction cosines of the planes on
which the shear strains are a minimum, that is, zero and also a maximum,
can be determined.
The following equations are obtained for l and m:
/[(13 1 - 13 3 W+ (13 2 -
13 3)m 2 - !(13 1 - 13 3)] = 0}
(3.44)
m[(13 1 - 13 3W+ (13 2 - 13 3)m 2 - !(13 2 - 13 3)] = 0
One solution to equations (3.44) is given when
l = m = 0 then n = ± 1.
when m = 0, l =I= 0 then l = ± (1/2) 112 from the first equation, and when
l = 0, m =F 0 then m = ± (1/2) 112 from the second equation.
58 ENGINEERING PLASTICITY

There are no general solutions for land m =I= 0 except when e1 = e2 .


If similar calculations are performed by eliminating first l and then m from
equation (3.43) the following table can be formed:

0 0 ±1 0 ± (1/2)1/2 ± (1/2)1/2
m 0 ±1 0 ± (1/2)1/2 0 + (1/2)1/2
n ±1 0 0 ± (1/2)1/2 ± (1/2)1/2 0

The first three columns give the direction cosines of the principal planes
on which the shear strains are zero. The other columns give directions for
angles of n/4. These planes consequently bisect the angles between the
principal planes and, on these planes, the shear strains have extremum
values. Substituting these values of the direction cosines into equation (3.43)
produces the shear strains y1 , y2 and y3 which are usually known as the
principal shear strains:

Y1: ±!(e2- E: 3}}


Y2- ± 2(e1 - e3) (3.45)
Y3 = ± !(e1 - e2)
If e1 ~ e2 ~ e3, the maximum shear strain acts on a plane bisecting the
angle between the maximum and minimum principal strain directions and
is equal in magnitude to half the difference of these strains, that is,
(3.46)

3.2.6 Octahedral strains


If the octahedral planes are now considered for which l = m = n = 1/(3) 1 12
so that they are inclined at equal angles of 54°44' to the principal directions,
these values of direction cosines can be substituted into equation (3.41) to
give the octahedral direct strain, that is, in a direction normal to an octahedral
plane,
(3.47)
The shear strain on an octahedral plane referred to as the octahedral
shear strain and designated, Yocr• is then given by substituting the particular
values of the direction cosines in equation (3.43).
Yocr 2 = !(ei + e~ + e~)- t(e1 + e2 + e3) 2
= H(e1 - e2) 2 + (e2 - e3) 2 + (e3 - e1) 2]
or (3.48)
It will be noted that the octahedral strains E:ocr and YocT given by equa-
STRAIN ANALYSIS 59
tions (3.47) and (3.48), respectively, are in exact analogy to the octahedral
normal and shear stresses.
In terms of the invariants of the strain tensor

YocT=-3-
(2)1/2 [J2
1+
31 J1/2
2 (3.49)

and in terms of non-principal strains,


YocT = H (ex- e/ + (ey- ez>l + (ez- exf + 6(e;Y + e;z + e;x) Jl12 (3.50)

3.2. 7 Representative or equivalent strain


The representative or equivalent strain is arbitrarily defined by
(2)1/2 [ ]1/2
€ = (2) 112 'l'ocT = -3- (el - e2) 2 + (e2 - e3) 2 + (e3 - e1f (3.51)

The representative or equivalent strain, e, cannot be visualised as acting


on a specified plane as in the case of the octahedral stresses and strains.
However, it has the property that in a state of strain defined by e2 = e3 =
1 -
- zel,e = el.

3.2.8 Spherical and deviator strains


As in the case of the stress tensor, the strain tensor, eij, can be resolveJ into
two parts: (a) a spherical or volumetric part, qij, which is proportional to the
volume change, and (b) a deviator part, e;P which represents a change in
geometry of the body and thus a pure distortional strain.
Changes in volume are usually considered negligible during plastic
deformation. The components of strain characterising change in shape
are therefore important.
The spherical strain tensor is given by
0
(3.52)

where em= t(e 1 + e2 + e3 } is the mean strain.


The deviator strain tensor then becomes

(3.53)

(3.54)
60 ENGINEERING PLASTICITY

or in terms of the principal strains,

By analogy with the deviator stress tensor discussed earlier the invariants of
the deviator strain tensor are

(3.56)

It also follows that


(3.57)

3.2.9 Strain rates and the strain rate tensor


From the finite strain tensor of equation (3.31) and referring to equation (3.12)
the direct strain of a line element can be defined as

Then, considering a small time interval bt during which a line element is


subjected to a small strain such that the coefficients of the finite strain tensor
are small compared with unity when the products of the derivatives in
equations (3.30) can be neglected and the components of displacement are
u,v and w,
. [ -&r
11m
<lt-+0bt
J derdt
=-=~:
.
r

Therefore . =~(au)p
er ox ot + +
oy ct m2 ~(ow)
~(av) oz ot n2 +
ay at ~(av)Jzm
[ ~(au)+ ax at +
[ ~(ov) ay ow)Jmn
az at + ~( at +

[ oxa ( at at J1n
aw) + oza(au) (3.59)

or (3.60)
STRAIN ANALYSIS 61

where

. ov
direct strain-rates oy
ey =

. ow
ez=az (3.61)
· ou ov
<Pxy =oy +ox
. ov ow
engineering shear strain-rates ¢yz = oz + oy
. ow ou
</Jzx =OX+ OZ
The quantities ex,iy,iz determine the rates of relative elongation or
contraction in the direction of the OX, OY and OZ axes, respectively.
The other quantities, <fixy• ~yz' ~zx determine the angular rates of change of
initially right angles.
The rate of volumetric or dilatational strain is then given by

= !(~:)+ :l~;)+ ~;) :l (3.62)

=div q
where q is the velocity vector.
Hence, analogous to the infinitesimal pure strain tensor of equation (3.9)
a strain-rate tensor can also be formed which is given by

ov =! [( ov; +
oxj
1
2 oxj ox, + ( oxj3)
OV; -~)]
OX;
(3.63)

= 811 + w11
the strain-rate tensor, 611 being the symmetric part. The skew-symmetric
part corresponds to a rigid body rotation of the element considered and a
necessary and sufficient condition for rigid body motion is that 611 = 0.
If w11 = 0 then the flow is irrotational. An extensive discussion of the strain-
rate and vorticity tensors is to be found in the textbook by Aris 1 .
Since velocities are the total derivatives of the displacement with respect
to time it is evident that
(3.64)

In the case of small deformation, simple relations exist between the strain
62 ENGINEERING PLASTICITY

components and the strain-rate components, namely,

v.=-u.
a
1 at 1 (3.65)

a
and eij =at eij (3.66)

[' Yuj
Yxy

Bij = ~yx ey ~yz (3.67)


Yzx Yzy ez
Besides the rate of pure shear strain characterised by the strain-rate tensor,
8ij, a volume element experiences rigid body displacement determined by
the translatory velocity q and a rotation with angular velocity
w=-!curlq (3.68)
In tensor notation, the components of the strain-rate are

ij
8 = ~{ a~j ( ~~i) + a~i ( ~)} (3.69)

=!(avi+3)
2 axj axi
where velocities vi = duidt are the components of the velocity vector q and
are with respect to the current position vector xi.
The velocity gradient tensor ovioxj may be written as the sum of sym-
metric and skew-symmetric parts.
The acceleration is given by

(3.70)

The translatory part in the expression for the total derivative is omitted
since for small strains it is usually possible to assume that the coordinate
derivatives of displacement and velocity are negligible.
It should also be stated that, in general,

(3.71)

since the principal axes of the strain tensor and the strain-rate tensor do not
coincide.

3.2.10 Increments in the strain components


The mechanical properties of metals, in conditions of relatively slow plastic
deformation at not too high a temperature, are known to be practically
STRAIN ANALYSIS 63
independent of the rate of deformation. In this case, the main interest then
lies not in the strain-rate but in the small strain increments eij dt denoted
d~:ii. It must be borne in mind, however, that generally speaking these
quantities are not differentials of the strain components.
These strain increments are determined in accordance with

d~:ii = ~ ( 8~i dui + 8~i dui) (3.72)

and they also generate a tensor and have a physical meaning. The relations
defined in equation (3. 72) are useful for describing large strains which may
be obtained by integrating the small changes. The increments in the strain
components are evaluated with respect to the instantaneous state.
If the principal axes do not rotate under deformation, then the integrals
fd~:ii have a simple physical meaning, being equal to the corresponding
logarithmic or natural strains. In this case, the strains are additive, that is,
the sum of successive natural strains is equal to the resultant natural strain.
However, in the general case, the integrals f d~:ii cannot be evaluated and
do not have a physical meaning. These integrals can be found only if the
strain path is known, that is, if the components d~:ii are known as functions
of some parameter such as, for example, the deforming force. This limits
the range of application of natural strains to the case of fixed principal
directions.

3.2.11 Logarithmic or natural strain


It was shown earlier that the dilatational or volumetric strain is given by
d = (1 + ex)(1 + ey)(1 + ez)- 1
or with the edges of the infinitesimal parallelepiped of figure 3.1 parallel to
the principal axes of strain this becomes
d = (1 + e 1 )(1 + e 2 )(1 + e 3 ) - 1
= e 1 + e2 + e 3 + e 1 e 2 + e 2 e 3 + e 3 e 1 + e 1 e 2 e 3
or
which for small strains, when products of strains can be neglected, equation
(3.11) shows that
d=Jl
Thus for large strains the first invariant of the strain tensor does not
represent the dilatational or volumetric strain rigorously. It is consequently
desirable to use a different measure of strain which is not affected by this
discrepancy. Ludwik 2 , in 1909, suggested the introduction of logarithmic
or natural strain.
If the deformed state is considered and the increment of strain is calculated
64 ENGINEERING PLASTICITY

according to the two definitions (a) for conventional or engineering strain,


and (b) for logarithmic or natural strain then denoting the initial length of
a line element by 10 and its deformed length by 1, the conventional or engineer-
ing strain is
e = (1- 10 )/10 = (1/1 0 ) - 1 (3.73)
The logarithmic or natural strain is
B = ln(1/1 0 ) = ln(1 +e) (3.74)
An increment d1 in the deformed length gives an engineering strain increment
as
de= d1/1 0 (3.75)
and the natural strain increment as
de= (d1/1 0 )(10 /1) = d1/1 = de/(1 +e) (3.76)
It follows that an increment of engineering strain expresses the change
in length with respect to the original length of the line element while incre-
ments of the natural strain are determined in terms of the instantaneous
length of the line element. For very small strains the two definitions of
strain and increment of strain give identical values (see figure 3.9).
Since B = ln(1 +e)
1 + e = exp(e)
or e = exp(e)- 1 (3. 77)

.!!!
c
...0
-;;; 1·0
...::>0
o0·5
z
l:l
c
~01---::-+:::---.~--+::-----:+::---
~ 1·5 2·0
c
·a
..
!; 0·5

o>
c

.
; 1·0
c
, _____ e= lnll/1 0 ) = tn( l+e)

o>
c
UJ

Figure 3.9. Engineering strain and natural strain


STRAIN ANALYSIS 65
Conventional dilatational strain in terms of principal strains is then given
by
.::\ = exp(e 1) exp(e 2 ) exp(e 3) - 1
(3.78)
= exp(e 1 + e2 + e3) - 1
and the natural dilatational strain
.::\(nat)= ln(1 + .::\) = e1 + ez + e3 = 11 (3.79)

3.2.12 Mohr circle diagram for incremental strains


The incremental direct strains, de, and the incremental shear strains, dy,
are analogous to the normal stresses, a, and shear stresses, -r, respectively,
in the stress equations. It has been previously shown that infinitesimal
strains and small finite strains or incremental strains have tensor character-
istics so that it is possible to represent incremental strains by means of a
Mohr circle diagram in a similar manner to stresses. The Mohr circle diagram
defining the state of strain at any point in a plastically deforming region for
plane strain deformation is shown in figure 3.10(a). Note that the direct
incremental strains dex and deY are assumed to be compressive and therefore
considered negative.
Assuming the non-principal values are known, then the general state of
plane strain is dex, deY, dyxy = dyyx. The radius of the circle is then given by

+d'Y y

I -d'Ymax
I
2(d&x+dEy)

(b)

Figure 3.10. Mohr circle diagram for incremental strains: (a) Mohr circle
diagram for plane strain; (b) physical diagram
66 ENGINEERING PLASTICITY

r2 = (dex- dey)2 d 2
2 + Yxy
= i{dex- de/ + dy;Y

Therefore r= H (dex- de/+ 4dy;Y }112 (3.80)


The principal incremental direct strain values are
del 1
d e3 = - 2(dex +dey)± r (3.81)

The incremental maximum shear strain directions are then determined


as S 1 = cp and S2 = cp +t anticlockwise from the OX direction and the
directions of the principal incremental direct strains are P 1 = cp + ~ and
P 2 = cp + 34n anticlockwise from the OX direction. These directions are
illustrated in the physical diagram of figure 3.10(b).

REFERENCES
1. Aris, R., Vectors, Tensors and Basic Equations of Fluid Mechanics,
Prentice-Hall, New Jersey (1962)
2. Ludwik, P., Elemente der technologischen Mechanik, Springer-Verlag,
Berlin, (1909)
4
Yield Criteria for Ductile Metals

When a body is subjected to a system of external forces the body is stressed


and it deforms. If the forces are relieved then the stresses are relieved and
the body may regain its original shape. In this case, the deformation is
recoverable or reversible and is referred to as elastic deformation. The associa-
ted stresses and strains are elastic stresses and strains.
However, the external forces may be of such magnitudes that, when
relieved, the deformation is not entirely recoverable and the body does
not regain its original shape. In this case, plastic, non-recoverable or irrever-
sible deformation has occurred. The stresses thus attain a value which exceeds
that required for elastic deformation and the material is said to have yielded.
A material is homogeneous if its properties do not vary from element
to element and the material is isotropic if its properties are independent
of the orientation of the system of coordinate reference axes chosen and
therefore independent of direction. The original workpiece material sub-
sequently used in a metal forming process may be regarded as being essen-
tially homogeneous and isotropic and it remains so during elastic
deformation up to the onset of yield. Nevertheless, during plastic deforma-
tion, the workpiece material will tend to become increasingly anisotropic
and inhomogeneity may be introduced.
The subject of this chapter is concerned with two yield criteria to predict
the stress state for the onset of plastic deformation of homogeneous and
isotropic ductile metals. A yield criterion for an anisotropic metal is also
considered.

4.1 GENERAL CONSIDERATIONS


A yield criterion may be defined as a hypothesis concerning the limit of
elastic deformation due to any possible stress state. By means of a yield
criterion it is then possible to decide whether plastic deformation takes
place or, indeed, is possible. Any proposed yield criterion should be verified
experimentally.
68 ENGINEERING PLASTICITY

If the material is isotropic, plastic yielding can then only depend on the
magnitudes of the three principal stresses and not on their directions. Any
yield criterion can thus be expressed in the form
(4.1)
where J 1, J 2, J 3 are the first three invariants of the stress tensor uii.
They have been previously defined in terms of the components of the
stress tensor, uii, in equations (2.28) but for convenience will be restated
here
J 1 = Ux + (1y + (1z

J 2 = - (uxuy + (1y(1z + uzux) + (t;y + r;z + r;x)


J3 = (1x(1y(1z + 2rxy 'l:yz 'l:zx- (r;zux + r;xuy + r;yuz)
In terms of the principal stresses which are the roots of the cubic equation
(13 - J 1(12 - J 2(1 - J3 = 0
the stress invariants are
J1 =0"1 +u2 +u3
J2= -(u1u2+u2u3+u3u1)
J3 = (11 (12(13
It has been shown experimentally by Bridgman 1, who performed tensile
tests on both metallic and nonmetallic materials subjected to very high
hydrostatic pressures of the order of 25 000 atm, that some materials are
compressible to a significant extent. For the range of pressures usually
encountered during metal forming processes the degree of compressibility
was found to be very small.
It may therefore be assumed that for a moderate hydrostatic stress, either
compressive or tensile, and whether applied alone or superimposed on a
combined stress state, the yielding of a metal is unaffected. Equation (4.1)
may therefore be simplified by stating the yield criterion in terms of the
invariants of the deviator stress tensor, u;i, so that
(4.2)
since J~ = 0 and
J~ = - (u~ u~ + u~u~ + u~u~) = { (u~) 2 + (u~) 2 + (u~) 2 }/2
J~ = ci~ u~ o-~ = { (u~) 3 + (u~) 3 + (u~) 3 } /3
An idealised plastically deforming body does not exhibit a Bauschinger
effect which implies that the magnitude of the yield stress is the same in
tension and compression. The Bauschinger effect is discussed in chapter 6.
Consider an element to be relieved from a plastic stress state, uii, and then
restressed to the state, - uii' keeping the ratio of the stress components
constant throughout. The condition assumed is that the element is deformed
YIELD CRITERIA FOR DUCTILE METALS 69
elastically and is about to yield. Since J~ changes sign with a reversal of
stress it follows that the function cjJ in equation (4.2) must be an even function
of J~ whilst a function of J~ satisfies the required condition.
The quadratic invariant of the deviator stress tensor, a;i, namely
J~ = -(a~ a~+ a~ a~+ a~ a~)= {(a'1 ) 2 + (a~) 2 + (a~) 2 } /2
can be restated as

However
2(a~ a~+ a~ a~+ a~a'1 ) = (a'1 +a~+ a~f- {(a'1 ) 2 + (a~f + (a~) 2 } (4.4)
Therefore
(a~f + (a~f + (a~) 2 =[(a~- a~) 2 +(a~- a~f +(a~- a'1 ) 2 +(a~+ a~+ a~) 2
- { (a~) 2 + (a~) 2 + (a~) 2 } ]/2 (4.5)
and since a~ + a~ + a~ = J~ = 0
3{(a~) 2 + (a~) 2 + (a~) 2 }/2 ={(a~- a~) 2 +(a~- a~) 2 +(a~- a~) 2 }/2
or (a'1 f + (a~) 2 + (a~f ={(a~- a~f +(a~- a~f +(a~- a'1 ) 2 }/3
Hence J~ ={(a~- a~) 2 +(a~- a~f +(a~- a~f}/6 (4.6)
If the values a~= al- am, a~= az- am, a~= a3- am are substituted
into equation (4.6) then J~ can be expressed in terms of the principal stresses
as

or in terms of the components of the stress tensor aii, as


J~ = [ {(ax- ayf + (ay- a 2)
2 + (az- aY} /6] + ('r:;Y + r;z + r;x) (4.8)

4.2 VON MISES YIELD CRITERION


It was suggested by von Mises 2 , in 1913, that the quadratic invariant of
the deviator stress tensor, J~, should be considered as a yield criterion.
Provided that J~ is less than a characteristic value of the material, k 2 , the
material does not yield and the deformation is elastic. If no strain hardening
occurs then J~ can never exceed the value k 2 when the material yields.
Hence J~ < k 2 during elastic deformation}
(4.9)
and J~ = k2 at yield
The von Mises yield criterion in terms of the components of the stress
tensor, aii, then becomes
70 ENGINEERING PLASTICITY

J~ = [ {(o-x- o-y + (o-y- O"z) 2 + (o-z- O"x) 2}/6] + (r;y + r;z + r;J = k 2 (4.10)
or in terms of the principal stresses
1~ = {(o-1- 0"2) 2 + (o-2- 0"3) 2 + (o-3- 0"1) 2}/6 = k 2 (4.11)
The characteristic value, k, of the material can be evaluated by means
of a uniaxial tensile test when the material is just yielding. Then o- 1 = Y
which is the uniaxial yield stress of the material, O" 2 = o- 3 = 0 and k is the
yield stress in pure shear.
Substituting these values into equation (4.11) produces
{(Y- Of+ 0 + (0- Yf} = 6k 2
that is, 2¥ 2 = 6k 2
or k = y /31/2 (4.12)
and the yield stress in pure shear is 1/3 1 / 2 times the yield stress in uniaxial
tension according to the von Mises yield criterion.
Equation (4.11) was also proposed independently by Huber 3 in 1904
and apparently by Maxwell in a letter to Kelvin as early as 1856. The von
Mises yield criterion was further interpreted by Hencky 4 to mean that
yielding commenced when the shear strain energy attained a critical value
corresponding to yielding in uniaxial tension.

4.3 TRESCA YIELD CRITERION


Tresca 5 suggested, in 1864, that yielding occurs when the maximum value
of the extremum shear stresses in the material attains a critical value. The
extremum shear stresses are given by equations (2.38) and the maximum
value is given by equation (2.39). The maximum value is equal in magnitude
to half the difference between the algebraic maximum and minimum principal
stresses.
It follows that the Tresca yield criterion requires the maximum and
minimum principal stresses to be known in advance. If, however, it can
be assumed that o- 1 ~ o- 2 ~ o- 3 then the value of the maximum shear stress,
"max' is given by equation (2.39).
For yielding in uniaxial tension when o- 1 = Y, o- 2 = o- 3 = 0
(4.13)
For pure shear, o- 1 = - o- 3 = k and o- 2 = 0. Substituting these values into
equation (4.13) produces
k- (- k) = y
or k= Y/2 (4.14)
That is, the yield stress in pure shear is half the yield stress in uniaxial
tension according to the Tresca yield criterion.
YIELD CRITERIA FOR DUCTILE METALS 71
Both the von Mises yield criterion and the Tresca yield criterion are
used in solving plastic metal forming problems. Both are independent
of hydrostatic stress and only depend on the components of the deviator
stress. The von Mises yield criterion is regarded as isotropic because each
of the nine components of the deviator stress has the same effect in the
yield expression.
However, the Tresca yield criterion is unaffected by the intermediate
principal stress u 2 • This implies that u 2 can vary between a maximum
value of u 2 = u 1 and a minimum value of u 2 = u 3 without affecting the
criterion expressed by J~ .
The von Mises yield criterion usually provides a better correlation, but
not always, with the experimental data for engineering metals than does
the Tresca yield criterion. The relative magnitudes of the principal stresses
must be known, a priori, for application of the Tresca yield criterion. If
the relative magnitudes of the principal stresses are known then the Tresca
yield criterion is easier to apply and leads to simplicity in mathematical
derivation.

4.4 YIELD SURFACE FOR AN ISOTROPIC PERFECTLY


PLASTIC MATERIAL
The yield criterion given by either equation (4.11) or equation (4.13) describes
a so-called yield surface. Provided the state of stress is enclosed by the
yield surface the condition of yielding is not attained and the deformation
is elastic. If the state of stress describes a point on the yield surface then
yielding is initiated. For an isotropic, perfectly plastic material, yielding
will continue and the yield surface does not change with plastic flow.
However, engineering metals do strain-harden when deformed at a tempera-
ture below their recrystallisation temperature and, in such cases, subsequent
yield surfaces are applicable after different histories of plastic deformation.
In the first instance, it is convenient to consider the yielding of an isotropic,
perfectly plastic metal according to the von Mises and Tresca yield criteria
corresponding to a biaxial stress state or plane stress state. Assuming u 3 = 0,
then equation (4.11) stating the von Mises yield criterion in terms of the
principal stresses reduces to
+ (u 2 - 0)2 + (0- u 1)2 = 6k 2 = 2Y 2
(u 1 - u 2 ) 2
2ui- 2u 1u 2 + 2u~ = 6k 2 = 2Y 2
Therefore
ui - u1 U2 + u~ = Y 2
or, in a dimensionless form,
(u 1 /Y) 2 - (u tfY)(u 2 /Y) + (u 2 /Y) 2 = 1 (4.15)
72 ENGINEERING PLASTICITY

which is the equation to an ellipse known as the von Mises ellipse when
plotted in the (J 1 (J 2 plane.
If the Tresca yield criterion is applicable, this asserts that yielding will
occur when any one of the following six conditions is attained
(J1 -(J2= ±Y
(J2- (J3 = ±y (4.16)
(J3- (J1 = ±y
and if (J 3 = 0 for the biaxial stress case
(J 1 - (J 2 = y if (J 1 > 0, (J 2 < 0
(J 1 - (J 2 = - y if (J 1 < 0, (J 2 > 0
(J2=Yif(J2>(J1>0

(J1=Yif(J1>(J2>0

(J1 = - Yif(J1 <(J2 <0


(J2=-Yif(J2<(J1<0 (4.17)
When this yield criterion is similarly plotted in the (J 1 (J 2 plane for the
biaxial stress state then a hexagon referred to as the Tresca hexagon results.
Each of the curves, that is, the von Mises ellipse and the Tresca hexagon,
is known as a yield locus when plotted in, for example, the (J 1 (J 2 plane for the
biaxial stress state and these loci are compared in figure 4.1 which is presented
in the dimensionless form.

von Misu ellipse


1o;1v 1~1o; /vHo;/vl+ 1<;/Yl~+l
.. _1

(o;-0)/Y =+I
Figure 4.1. Yield loci for biaxial stress states according to the von Mises and
Tresca yield criteria
YIELD CRITERIA FOR DUCTILE METALS 73
4.5 HAIGH-WESTERGAARD STRESS SPACE
REPRESENTATION OF YIELD CRITERIA
In section 4.4, the yield surface and the two-dimensional yield loci according
to the von Mises and Tresca yield criteria for a biaxial stress state were
discussed. For the more general case, however, the yield criterion will be
a function of the nine components of the stress tensor, aij, which can be
reduced to the six independent components of the stress tensor. If the
material is assumed to be isotropic so that yielding is not affected by rotation
of the axes of reference then the principal stress axes can be chosen as the
references axes and the initial yield condition can be stated as
(4.18)
Since it has already been assumed that hydrostatic stress states do not
influence yielding and thus only the deviator stresses are involved, the yield
function can be restated as
(4.19)
or, alternatively, in terms of the invariants of the deviator stress tensor,

Hydrostatic
stress states
Vector sum of o;-
CJ2-CJ3
deviator


0"3 stress components

c . . . . ...
... ..-"1'-.., (a1,a;zP"3 l
...
... ..... p
"'
//

Vector sum of
hydrostatic
stress
components

0:I
11-plane
o;+o;+a3=o
Figute 4.2. Haigh-Westergaard stress space representation of yield criterion
74 ENGINEERING PLASTICITY

(4.20)
since J~ = 0.
The yield function has therefore been simplified to a function of the two
non-vanishing invariants of the deviator stress tensor and this function
is symmetric in the principal stresses. Whatever yield function is adopted
it must be symmetric in the principal stresses.
In figure 4.2 are shown three mutually orthogonal axes introducing the
(u 1 , u 2 , u 3 ) coordinate system which represents a three-dimensional stress
space called the Haigh-Westergaard stress space6 •7 • If the principal stresses
at a point on a body are (u 1 , u 2 , u 3 ) then this stress state is represented
by the bound vector 0 P in this stress space and the coordinates of P are
u 1> u 2 and u 3 •
The stress state may therefore be written as the sum of the three vectors,
OA = upOB = u 2 and OC = u 3
Therefore IOP I= (u~ + u~ + uW 12 (4.21)
Consider an axis OE passing through the origin which is equally inclined
to the three principal stress axes then its direction cosines are l = m = n =
1/3 1 ' 2 and the equal angle of inclination of the axis OE to the three stress
axes is cos- 1 (1/3) 1' 2 = 54°44'. For every point on this axis the stress state
is one where
(4.22)
which corresponds to a hydrostatic stress state and the deviator stresses
are equal to zero.
The equation to any plane which is perpendicular to OE will be
(4.23)
where d is the distance along the normal OE from the origin to the plane.
The hydrostatic or spherical component of the stress tensor therefore
increases in a linear manner with the distance of the plane from the origin.
For the plane passing through the origin, the hydrostatic stress is zero and
u 1 + u 2 + u 3 = 0. This plane is known as then-plane or synoptic plane.
Any arbitrary stress state such as that defmed by the vector 0 P in figure
4.2 can therefore be resolved into two components, namely, the component
ON in the direction of OE and the component NP which is perpendicular
to ON and parallel to the n-plane.

ION I = (1/3)1/20" 1 + (1/3)1/20" 2 + (1/3)1/20" 3


= (1/3)112(0"1 + 0"2 + 0"3)
= 31/20"m (4.24)
Also
YIELD CRITERIA FOR DUCTILE METALS 75
= (ui + 0"~ + u~)- 3um
= (u 1 - um) 2 + (u 2 - um) 2 + (u 3 - umf
= (u~)2 + (u~f + (o-~)2
Therefore IPNI2 = 2J~ (4.25)
and, consequently, the components of PN are the deviator stresses, u~,
u~ , and u~ .
If a different stress state defined by a point P 1 is considered which is on
a line passing through P and parallel to OE then the projection of the vector
OP 1 on to then-plane is the same as that for the vector OP. The two states
of stress corresponding to the points P and P 1 therefore have the same
deviator stress components and only differ in their hydrostatic stress compo-
nents. All points on a line through P parallel to OE will thus have the same
deviator stress components. Since it is assumed that yielding is only deter-
mined by the deviator stress state and is not influenced by the hydrostatic
stress component, it follows that if one of the points such as P or P 1 on a
line parallel to OE lies on the yield surface then all such points will lie on
the yield surface. In this case, the yield surface will be a cylinder with genera-
tors which are parallel to OE. The intersection of the yield cylinder with
any plane perpendicular to its axis produces a curve which is a yield locus.
This yield locus is the same for all planes which are perpendicular to the
axis of the yield cylinder. It is therefore most convenient to choose the
n-plane on which the hydrostatic stress state is zero. Since this plane is
inclined to the principal stress axes at equal angles, their projections in this

Yield locus


-0'3

Figure 4.3. A possible yield locus for an isotropic, perfectly plastic metal shown
in the n-plane
76 ENGINEERING PLASTICITY

E
n- plane

(a)

(b)

Figure 4.4. (a) Projection of the principal stress component a 1 on to then-plane;


(b) polar coordinates of the point P on then-plane

plane must be inclined at 120o to each other as shown in figure 4.3 and the
yield locus must be symmetrical and the same shape in each of the six 60°
sectors dividing the n-plane.
Let the polar coordinates of the point P in the n-plane be (r, 0) and a
and b the horizontal and vertical components of r, as shown in figure 4.3.
The components of 0 P in the principal stress space are a 1> a 2 and a 3 •
Considering, for example, the projection of the component a 1 on to the
n-plane then by referring to figure 4.4(a) it can be seen that this is equal
to {ai - (1/3)ai} 112 = (2/3) 112 a 1. The projections of a 2 and a 3 on to the
n-plane can be obtained in a similar manner so that the projections of the
three principal stress components on to the n-plane are (2/3) 1i 2a 1> (2/3) 1i 2a 2
and (2/3) 1i 2 a 3 , respectively.
From figure 4.4(b) the following relationships can now be deduced
a= (2/3) 1i 2a 2 cos 30° - (2/3) 1i 2a 1 cos 30°
= (2/3)1f2a 2(31/2 /2) - (2/3)1f2a 1(31/2 /2)
YIELD CRITERIA FOR DUCTILE METALS 77
Therefore a=(a 2 -a 1)/2 112 (4.26)
b = (2/3) 1' 2a 3 - (2/3) 1' 2a 2 sin 30° - (2/3) 1' 2a 1 sin 30°
= (2/3)1/20"3- (1/6)1/20"2- (1/6)1/20"1
Therefore b = (1/6) 112 (2a 3 - a 2 - a 1) (4.27)
r 2 = a 2 + b 2 = (a 2 - a 1)2/2 + (1/6)(2a 3 - a 2 - a 1f
= (1/3){(a1- az)2 + (a2- a3)2 + (a3- a1)2}
= (a 1 - am) 2 + (a 2 - am) 2 + (a 3 - amf
= (a~) 2 + (a~f + (a~) 2
Therefore r2 = 21~ (4.28)
e= tan- 1(bja) = tan- 1[(2a3- 0"2- a1)/{3 112(a2- a1)}]
or 3 112 tan e= (2a3- 0"2- a1)/(az- a1) (4.29)

4.5.1 Representation of von Mises and Tresca yield criteria on the


n- plane and in principal stress space
The von Mises yield criterion in terms of the principal stresses is defined
by equation (4.11)
1~ = kz = Y2/3
and from equation (4.28) it follows that
r2 = 21~ = 2Y 2 /3 (4.30)

. . 0'3
von M1su Circle

Trcsca hexagon

Figure 4.5. Representation of the von Mises and Tresca yield criteria on the
n-plane
78 ENGINEERING PLASTICITY

The yield locus is therefore a circle of radius r = (2/3) 1' 2¥, as illustrated
in figure 4.5, and the yield surface in the Haigh-Westergaard stress space
is a circular cylinder of radius (2/3) 1' 2Y with its geometric axis OE passing
through the origin and equally inclined to the three principal stress axes.
The stress state represented by any point in the sector between the a 2
and - a 1 axes in figure 4.5 is such that

and the Tresca criterion applicable to yielding for this sector is then given
by a 2 - a 1 = Y. The Tresca yield criterion for this sector is thus represented
by a straight line which is parallel to the a 3 axis, since it is independent of
a 3, and is at a distance equal to (a 2 - a 1)/(2) 112 = Y/(2) 112 from it, as given
by equation (4.26). In a similar manner, a straight line is obtained for each
sector to produce the complete yield locus for the Tresca yield criterion
which is a regular hexagon. This is also illustrated in figure 4.5 for purposes
of comparison with the yield locus for the von Mises yield criterion. It
will be noted that the von Mises circle passes through the comers of the
Tresca hexagon since the radius to a comer ofthe hexagon is given by
r = (Y /2 112 )/cos 30° = (2/3) 1' 2¥
which is also the radius of the von Mises circle. From equation (4.29) when
(} = oo
2a3 -a2 -a 1 =0
or a 3 = (a 1 + a 2)/2
and the hydrostatic stress
am= (a1 + a 2 + a 3)/3 = (a 1 + a 2)/2
If the hydrostatic stress is now subtracted from the stress components
a 1, a 2 and a 3 the deviator stress state is given by a~ = ( a 1 - a 2)/2, a~ =
(a 2 - a 1)/2, a~= 0 which defines a state of pure shear. It is then evident
from figure 4.5 that the yield loci for the von Mises and Tresca yield criteria
differ most for the case of pure shear when the von Mises yield criterion
predicts a yield stress 2/3 1' 2 times that given by the Tresca yield criterion.
When(}= 30°, tan(}= 1/3 1' 2 and from equation (4.29)
2a3- a2- a1 = a2- a1
or 2a 3 - 2a 2 = 0
Therefore
and if the hydrostatic stress of a 2 is subtracted from the principal stress
components a 1, a 2 and a 3 the deviator stress state is a~ = ( a 1 - a 2), a~ = 0,
a~ = 0 which corresponds to a uniaxial stress state.
In the Haigh-Westergaard stress space, the Tresca yield surface is a
YIELD CRITERIA FOR DUCTILE METALS 79
E

von Misu
circular cylinder

Figure 4.6. The yield surfaces for the von Mises and Tresca yield criteria in the
Haigh-Westergaard stress space

regular hexagonal cylinder which is inscribed within the von Mises circular
cylinder as shown in Figure 4.6.

4.6 EXPERIMENTAL VERIFICATION OF YIELD CRITERIA


The most common type of test specimen which has been used to experi-
mentally investigate the yield criteria is a thin-walled tube subjected to
combined stress. The combined stress state can be achieved, for example,
by subjecting the tube simultaneously to a couple C, an axial force F and
an internal hydrostatic pressure p. By varying these parameters it is then
possible to obtain different stress combinations which result in different
magnitudes of principal stresses and different principal stress directions.
However, assume the thin-walled tube shown in figure 4.7(a) to be subjected
to a couple C, which produces elastic deformation, and then a tensile axial
force F to be applied so as to just cause yielding. The angle of twist and the
varying values of axial extension of the tube are noted as the tensile axial
force increases. Yielding is assumed to have occurred when the axial strain
noticeably increases as discerned from an equivalent true stress, 0:-equivalent
natural strain, e, characteristic curve for the test material of the type presented
in figure 4.7(b).
For an element of the tube wall at the point P, the tensile axial stress is
a and the shear stress is -r. Since the tube is thin walled the shear stress
distribution across the wall can be assumed to be sensibly constant so that
the shear stress -r is constant.
80 ENGINEERING PLASTICITY

F F

(a)

'i ~lcld polot

..
!:; I

.."...
:::J

c
">
0
·;
CT
1&.1.___ _ _ _ _ _ _ _ __

Equivalent natural strain E


(b)
Figure 4.7. (a) Thin-walled tube subjected to a pure couple C and an axial
force F showing the stress state at a point P; (b) equivalent true stress, a--equivalent
natural strain, echaracteristic curve for the test material

The principal stresses at the point P in the tube wall at any instant are
u 1 = (u/2) + {(u2 /4) + 'r2}1/2
u2 = (u/2)- {(u2/4) + 'r2}1/2
(J3 =0 (4.31)
Therefore ul- u2 = (u2 + 4'r2)1i2 (4.32)
and (u 1 - u 2)2 + (u 2 - u 3 ) 2 + (u 3 - u 1)2 = 2u 2 + 6"C 2 (4.33)
By using equation (4.32) the Tresca yield criterion predicts
(u /Y) 2 + 4('r/Y) 2 = 1 4.34)
and qsing equation (4.33) the von Mises yield criterion predicts
2o-2 + 6"C2 = 2Y2
or (u /Y) 2 + 3('r/Y) 2 = 1 (4.35)
where Y is the uniaxial yield stress of the tube metal in tension.
YIELD CRITERIA FOR DUCTILE METALS 81
~6r-----~------,-------.------.------.

von Misu yield criterion

0-4

'f/Y
Trcsca yield criterion
0·2

0 02 0•4 08 1·0
CJ/Y

Figure 4.8. The Tresca and von Mises yield criteria represented in the dimension-
less (u I Y)-( t I Y) plane for a biaxial stress state
Both equations (4.34) and (4.35) plot as ellipses in the ur: plane and in
the dimensionless form presented in figure 4.8. Experimental data obtained
in this manner are found to plot between the two ellipses although generally
closer to the von Mises ellipse. These deviations from the theoretical predic-
tions are usually partially attributed to a degree of anisotropy of the specimen
material and to experimental inaccuracy.
The classical experimental work of this kind was performed in 1931
by Taylor and Quinney 8 which was intended to solve this problem. They
used copper, mild steel and aluminium thin-walled tubes, which were said
to be very nearly isotropic, and tested them in combined torsion and tension.
However, they also observed similar deviations from the theoretical curves
and concluded that these discrepancies were real and could not be attributed
to experimental error or anisotropy of the specimen material.
Similar results were obtained earlier, for example, by Lode9 in 1926,
by Ros and Eichinger 10 in 1929 and, in 1953, by Siebel 11 who employed
combined bending and torsion. Details of other tests of a similar nature
can be found in the literature and an extensive survey of the literature
concerned with this topic up to about 1948 was presented by Drucker 12 .
A different method of producing a combined stress state to experimentally
investigate yield criteria has been employed more recently. This method
follows the theory of localised oblique 'necking' in a thin metal strip as
advanced by Hi11 13 •14• A thin uniform rectangular section metal strip in
which a groove is machined and where the width of the strip is not less
than five times the thickness is subjected to uniaxial tension. This method
was used by Lianis and Ford 15 in 1957 and by Parker and Bassett 16 in
1964. Attempts have also been made by Stockton and Drucker 17 to improve
the correlation of experimental data with the yield criteria by introducing
the third invariant, J~, of the deviator stress tensor into the yield criterion.
82 ENGINEERING PLASTICITY

Some degree of anisotropy is likely to be initially present in the material


to be tested and this probably becomes more pronounced as deformation
proceeds. Together with the inevitable experimental inaccuracies, these
factors will certainly affect the correlation of experimental data with, for
example, the von Mises yield criterion. However, an important source of
deviation is most likely to be due to the real difficulty in defming the magni-
tude of the yield stress experimentally. This point is further discussed in
chapter 6. Furthermore, the composition and properties of an engineering
metal which has the same nominal specification are known to vary, to
some extent, throughout the metal and also from batch to batch during
manufacture. It therefore seems that the von Mises yield criterion predicts
yielding for most engineering metals with sufficient accuracy. There is no
great advantage to be gained from any yield criterion based on a theory
which is mathematically more exact, particularly since it will most certainly
be more complex.

4.6.1 The Lode stress parameter


It was suggested in the previous section that Lode was responsible for the
first experiments using a thin-walled tube to investigate the validity of the
yield criteria. He used tubes of steel, copper and nickel subjected to various
combinations of uniaxial tension and internal hydrostatic pressure and
also devised a sensitive method to determine the effect of the intermediate
principal stress on yielding.
Assuming u 1 ~ u 2 ~ u 3 , then according to the Tresca yield criterion,
u 1 - u 3 = Y and therefore
(4.36)
The intermediate principal stress, u 2 , can thus vary from a maximum
value u 2 = u 1 to a minimum value u 2 = u 3 without apparently affecting the
yield criterion expressed by equation (4.36). To characterise the influence
of the intermediate principal stress, u 2 , in the von Mises yield criterion,
Lode introduced the parameter
= (2u 2 - u 3 - u 1 )/(u 1 - u 3 )
J1.
f1.=u 2 -{(u 1 +u 3 )/2}/{(u 1 -u 3 )/2} (4.37)
which is known as the Lode stress parameter and corresponds with equation
(4.29)when
J1. = - 3112 tan (} (4.38)
and the stress state is represented by any point in the sector of the n-plane
between the u 1 and - u 3 axes in figure 4.5 such that u 1 ~ u 2 ~ u 3 • Equation
(4.37) can be rearranged so that
(4.39)
YIELD CRITERIA FOR DUCTILE METALS 83
The von Mises yield criterion in terms of the principal stresses is
(0"1- 0"2f + (0"2 + 0"3) 2 + (0" 3 - 0" 1) 2 = 6k 2 = 2Y 2
and if 0" 2 from equation (4.39) is substituted then, after rearranging and
simplifying, the von Mises yield criterion becomes
(4.40)
where f1 is the Lode stress parameter defined by equation (4.37).
When O" 2 = O" 3 , equation (4.37) shows that f1 = - 1 and when O" 2 = O" 1 ,
then f1 = + 1. Since O" 1 ;?; O" 2 ;?; O" 3 it follows that - 1 ~ f1 ~ + 1.
When f1 = 0, 20" 2 - O" 3 - O" 1 = 0 and O" 2 = (O" 1 + O" 3 )/2. The principal stresses
are then O" ~' O" 2 = (O" 1 + O" 3 )/2, O" 3 . The hydrostatic stress O"m = (O" 1 + O" 2 +
O" 3 )/3 = (O" 1 + O" 3 )/2 and the deviator stresses are O"~ = (O" 1 - O" 3 )/2, O"~ = 0
and O"~ = (O" 3 - O" 1 )/2. This is a state of pure shear and is represented on the
n-plane by(}= 0°.
When f1 = - 1, the principal stresses are O" 1 , O" 2 = O" 3 which is a uniaxial
tension (O" 1 - O" 3 ) with a hydrostatic stress O" 3 and is represented on the
n-plane by(}=+ 30°. When f1 = + 1, the principal stresses are 0" 1 = 0" 2 , 0" 3
which is a uniaxial compression (0" 3 - 0" 1 ) with a hydrostatic stress 0" 1 and
is represented on the n-plane by (} = - 30°.
The difference between the Tresca and von Mises yield criteria is indicated
by the extent to which the right-hand side of equation (4.40) differs from
unity. The two yield criteria are expected to agree when f1 = ± 1. Lode
conducted his tests for the range (- 1 < f1 < + 1) corresponding to ( + 30° >
(} > - 30°) so that all combinations of stress state from uniaxial tension,
including pure shear to uniaxial compression, were involved.
Equation (4.40) is presented in figure 4.9 excluding Lode's experimental
data which indicated that the actual yield stress values occurred between
the two predictions although generally closer to the von Mises yield criterion.

1·25r-----.------l-------,,...--------,
2/13 von Miscs

-----
1·20 Trcsca yield criterion yield criterion 1
1o;- CJ3 )/Y•I lo;-c>;l/Y =2/13+)h2
HS

1-10

1-QS

-0·5 0 +C>S
Lode parameter p
Figure 4.9. Lode's comparison ofthe yield criteria (experimental results excluded)
84 ENGINEERING PLASTICITY

The maximum difference between the two yield criteria occurs when Jl = 0
which corresponds with the case of pure shear and is 2/3 112 , which agrees
with that stated earlier. Lode's stress parameter, Jl, and a plastic strain
parameter are discussed further in section 5.6.

4.7 SUBSEQUENT YIELD SURFACES FOR AN ISOTROPIC


STRAIN-HARDENING MATERIAL
When considering the yield surface in stress space and the yield locus on
the n-plane appropriate to either the Tresca or the von Mises yield criterion
it was previously assumed that the material was isotropic and perfectly
plastic. This implied that the yield surface remained unchanged and the
stress after yielding remained constant as plastic deformation proceeded.
Therefore, if yielding occurred such that the von Mises yield criterion was
applicable then the yield locus on the n-plane was a circle and the yield
surface in stress space was a circular cylinder. However, for a strain-hardening
material, where the stress required to maintain plastic flow increases with
increasing plastic strain, the yield surface must change if straining continues
beyond the initial yield condition.
Consequently, if the initial yield stress is denoted by Y0 , the initial yield
locus on then-plane is a circle of radius (2/3) 112 Y0 .If straining now continues
beyond the initial yield condition to a value such that the stress required
is, say, Y., and then this stress is relieved, upon reloading, the yielding would
not occur again until the stress Y. is attained. Assuming that the further
plastic deformation did not introduce anisotropy of the material then the
subsequent yield locus is also a circle of radius (2/3) 112 Y•. On the n-plane,
this subsequent yield locus surrounds the initial yield locus and is concentric
Subsequent yield locus

Initial yield locus

(a) (b)
Figure 4.10. Subsequent yield locus on then-plane for an isotropic, strain-harden-
ing material according to (a) the von Mises yield criterion, and (b) the Tresca yield
criterion
YIELD CRITERIA FOR DUCTILE METALS 85
with it. The yield surface must therefore move outwards in some way, at
least, at the point where initial yield occurred. If the material remains
isottopic and is strain-hardening then the circular yield locus expands with
stress and strain history. The yield locus for an isotropic strain-hardening
material which deforms according to the Tresca yield criterion appears
on the n-plane as a series of concentric regular hexagons. These subsequent
yield loci for the two yield criteria are illustrated in figure 4.10.
For a perfectly plastic material, a yield function can be defined by the
relation
(4.41)
so that when the function 4J becomes equal to the constant K, yielding is
initiated and is represented by an initial yield surface in the Haigh-Wester-
gaard stress space or an initial yield locus on the n-plane. However, for
a strain-hardening material the value of K will alter depending on the
strain-hardening properties of the material. It is then convenient to consider
the function 4J as a loading function which represents the application of
the stress and the function K as a strain-hardening yield function which
depends on the previous stress and strain history of the material and also
on its strain-hardening properties.
Three separate cases for a strain-hardening material can then be identified:

(1) When 4J = K, the stress state is represented by a point on the yield


surface and if
d(jJ = (84J/8uij)·du;j > 0
this constitutes loading and indicates that the stress state is moving outwards
from the initial yield surface and plastic flow occurs.
(2) When <P = K and d</J = 0 this condition corresponds to the case where
the stress state remains on the initial yield surface, which implies for a
strain-hardening material that no plastic flow occurs. This condition is
referred to as neutral loading.
(3) When 4J = K and d(jJ < 0 this indicates that the stress state is moving
inwards from the initial yield surface and elastic unloading takes place.

The condition when 4J < K thus represents an elastic stress state. For a
perfectly plastic material 4J = K, dF = 0 for plastic flow and the case d(jJ > 0
is inapplicable.
The concept of an isotropic, strain-hardening material is mathematically
the simplest one. However, it may be considered only as a first approximation
because it does not take the Bauschinger effect into account. This effect
is known to contract the yield locus on one side whilst that on the other
side is expanded. The yield surface therefore changes shape as plastic defor-
mation progresses. The Bauschinger effect was verified in 1958 by experiments
performed by Naghdi, Essenburg and Koff18 • These investigators carried
86 ENGINEERING PLASTICITY

s~cond subs~qu~nt
yi~ld locus

+1"

l
or------------------------------H
-+Cf
First subs~qu~nt yi~ld locus
s~cond subs~qu~nt
yi~ld locus
_.,.

Figure 4.11. Diagrammatic representation of initial and subsequent yield loci


for a strain-hardening metal showing the Bauschinger effect

out tests with aluminium alloy tubes, initially with axial tension only,
and then with various ratios of torsion and axial tension to obtain an initial
yield locus. By unloading and reloading in a specified manner, subsequent
yield loci were also obtained. The results of these experiments are presented
diagrammatically in the ar plane of figure 4.11. By referring to this figure
it is evident that the initial von Mises elliptical yield locus does not expand
symmetrically due to the presence of the Bauschinger effect which is reflected,
for example, in the progressively reduced yield stresses required during
reversed torsion. This, of course, indicates that the concept of an isotropic,
strain-hardening material must be regarded only as a first approximation
for the plastic deformation of real engineering metals.

4.8 A YIELD CRITERION FOR AN ANISOTROPIC MATERIAL


The application of the Tresca and von Mises yield criteria has been shown
to be limited essentially to isotropic materials even though the theory can
be extended for strain-hardening provided the Bauschinger effect is absent.
Nevertheless, when an initially isotropic metal is subjected to plastic defor-
YIELD CRITERIA FOR DUCTILE METALS 87
mation, the crystallographic axes tend to rotate to a preferred orientation
because of the nature of the glide mechanism. The metal thus tends to
become increasingly anisotropic during plastic deformation. All engineering
metals exhibit anisotropy to some extent when plastically deformed at
ambient temperature such that, for example, the stress-strain relationship
for the metal varies in different directions. The degree of anisotropy and
its nature is known to depend on the forming processes and heat treatments
to which the metal may have been previously subjected. The effects of
anisotropy on the deformation of workpiece materials during metal forming
processes such as forging, extrusion and drawing are not very significant.
However, in the case of sheet metal forming, such as the manufacture of
stainless steel domestic sink units, the practical difficulties and the economic
implications are of considerable importance. For these reasons, the main
interest in the study of the effects of anisotropy during metal forming pro-
cesses has been concentrated on sheet metal and this topic is discussed
further in chapter 6.
A theory for the anisotropic behaviour of a sheet metal was proposed
by Jackson, Smith and Lankford 19 in 1948 and Dorn 20 in 1949. However,
the theory proposed by Hill 21 is presented here. This assumes the relatively
simple case of orthotropic anisotropy, that is, there are three mutually
orthogonal planes of symmetry at every point and the intersections of
these planes are considered as the principal axes of anisotropy. The axes
may vary in direction throughout the deforming material and even at a
given element during deformation as is the case for pure shear. Assuming
a rolled sheet metal, then the principal axes of anisotropy at a point are
in the direction of rolling of the sheet, transversely in the plane of the sheet
and in the thickness direction normal to this plane.
Because the von Mises yield criterion is known to approximately describe
the yielding of an isotropic, strain-hardening material, Hill suggested that
the simplest yield criterion for an anisotropic material would be one which
reduced to the von Mises yield criterion when the anisotropy became
negligible. It is further assumed that the Bauschinger effect is absent. The
yield criterion for an anisotropic material proposed by Hill in terms of
the components of the stress tensor, O";i, and with the principal axes of
anisotropy coinciding with cartesian axes of reference is given by
2f(a;j) =F(ay- az) 2 + G(az- ax) 2 + H(ax- al + 2L'r:;z + 2Mr;x + 2Nr;Y = 1
(4.42)
where F, G, H, L, M and N are parameters which are characteristic of the
current state of anisotropy. Linear terms. are excluded because it is assumed
that the Bauschinger effect is absent and that hydrostatic stress states do
not affect yielding.
Assuming the tensile yield stresses in the principal directions of anisotropy
to be X, Y and Z, respectively, then it can be shown that the parameters
F, G and Hare related to the yield stresses in the following manner:
88 ENGINEERING PLASTICITY

1/X 2 = G+H
1/¥ 2 =F+H
1/Z2 =F + G
(4.43)
2F = (1/¥ 2 ) + (1/Z 2 ) - (1/X 2 )
2G = (1/Z 2 ) + (1/X 2 ) - (1/¥ 2 )
2H = (1/X 2 ) + (1/¥ 2 ) - {l/Z 2 )
The condition for the anisotropy to be rotationally symmetric about
the z-axis corresponding to planar isotropy in the xy-plane implies that
the coefficients in equation (4.42) must be invariant and then it can be
shown that
N=F+2H=G+2H
and L=M (4.44)

For complete spherical symmetry, that is, complete isotropy


L=M=N= 3F= 3G= 3H (4.45)
and if these particular values of the parameters are substituted into equation
(4.42) the yield criterion reduces to
(ay- az) 2 + (az- aJ 2 +(ax- a/+ 6(r;z + r;x + r;y) = 6k 2 = 1/F (4.46)

REFERENCES
1. Bridgman, P. W., Studies in Large Plastic Flow and Fracture with
Special Emphasis on the Effects of Hydrostatic Pressure, McGraw-Hill,
New York (1952)
2. von Mises, R., Mechanik der festen Korper im plastisch deformablen
Zustand, Nachr. Ges. Wiss. Gottingen, 582 (1913)
3. Huber, M. T., Czasopismo techniczne, 22, 181, Lemberg (1904)
4. Hencky, H. Z., Zur Theorie Plasticher Deformationen und der hierdurch
im Material hervorgerufenen Nachspannungen, Z. angew. Math.
Mech., 4, 323 (1924)
5. Tresca, H., Sur l'ecoulement des corps solides soumis a
de fortes
pressions, C. R. Acad. Sci. Paris, 59, 754 (1864)
6. Haigh, B. P., Elastic limit ofa ductile metal, Engineering, 10, 158 (1920)
7. Westergaard, H. M., On the resistance of ductile materials to combined
stresses, J. Franklin Inst., 189, 627 (1920)
8. Taylor, G. I. and Quinney, H., The plastic distortion of metals, Phil.
Trans. R. Soc., A230, 323 (1931)
9. Lode, W., Versuche tiber den Einfluss der mittleren Hauptspannung auf
das Fliessen der Metalle Eisen, Kupfer und Nickel, Z. Phys., 36,
913 (1926)
YIELD CRITERIA FOR DUCTILE METALS 89
10. Ros, M. and Eichinger, A., Versuche Zur Klaerung der Frage der
Bruchgefahr III, Metalle, Eidgenoss. Material pruf. und Versuchsanstalt
Industriell Bauwerk und Gewerbe, Diskuss Ber. no. 34, Zurich, p. 3
(1929)
11. Siebel, M. P. L., The combined bending and twisting of thin cylinders
in the plastic range, J. Mech. Phys. Solids, 1, 189 (1953)
12. Drucker, D. C., Stress-strain relations in the plastic range-a survey of
theory and experiment, O.N.R. Report, NR-041-032 (1950)
13. Hill, R., The Mathematical Theory of Plasticity, ch. 12, p. 323, Oxford
University Press, London (1950)
14. Hill, R., On discontinuous plastic states with special reference to
localised necking in thin sheets, J. Mech. Phys. Solids, 1, 19 (1953)
15. Lianis, G. and Ford, H., An experimental investigation of the yield
criterion and the stress-strain law, J. Mech. Phys. Solids, 5, 215 (1957)
16. Parker, J. and Bassett, M. B., Plastic strain relationships-Some
experiments to derive subsequent yield surface, Trans. Am. Soc. mech
Engrs., SerE, 31, 676 (1964)
17. Stockton, F. D. and Drucker, D. C., Fitting mathematical theories of
plasticity to experimental results, J. Col!. Sci. (Rheology Issue), 5,
239 (1950)
18. Naghdi, P. M., Essenburg, F. and Koff, W., An experimental study of
initial and subsequent yield surfaces in plasticity, Trans. Am. Soc. mech
Engrs, 80; J. appl. Mechs., 25, 201 (1958)
19. Jackson, L. R., Smith, K. F. and Lankford, W. T., Plastic flow in
anisotropic sheet metal, Metals Technology Tech. Pub. No. 2440 (1948);
J. Metals, 1, 323 (1949)
20. Darn, J. E., Stress strain relations for anisotropic plastic flow, J. appl.
Phys., 20, p. 15 (1949)
21. Hill, R., The Mathematical Theory of Plasticity, ch. 12, p. 317, Oxford
University Press, London (1950)
5
Stress-Strain Relations

In chapter 2, the state of stress at a point in a deforming body was considered


and chapter 3 dealt separately with the state of strain. The analyses did
not involve the properties of the deforming material. It follows that all
the derived equations are generalised and apply to any material although
the primary consideration here will be confmed to engineering metals. Two
yield criteria for ductile metals were developed in chapter 4 so that it is
possible to estimate whether the deformation is elastic or plastic.
For the formulation of a plasticity theory it now remains to relate the
stress tensor to the strain tensor. A theory is required to be concerned
with the behaviour of the metal which may or may not completely correlate
with experimental observation. The problem of relating the stress tensor
with the strain tensor is the subject of this chapter. However, the reader
is referred to a more extensive treatment of this subject by Hill 1 which
presents an elegant account of the foundations ofthe theory.
It will be assumed that the deforming metal is perfectly elastic when
stressed to a state without the yield condition being attained. The metal is
assumed to be isotropic and homogeneous. During elastic deformation, the
elastic constants are assumed not to vary with prestrain and are independent
of the stress and strain paths. The hysteresis loop which may be encountered
during loading and unloading and the Bauschinger effect are neglected. It
is then implied that the elastic strain state is a unique function of the elastic
stress state pertaining at a given instant and is independent of how that
stress state was attained. This is not the case when the metal has yielded
and plastic deformation occurs. The strain state is then dependent on the
stress history and the stress-strain relation is generally nonlinear. It is
therefore necessary when considering the plastic deformation of a metal
to relate the stress to the strain increment and the final strain is then obtained
by integrating the strain increments as determined by the stress history.
Simplified stress-strain relations for an anisotropic material are also
developed.
It should be remembered that the shear strain components of the symmetric
STRESS-STRAIN RELATIONS 91
part of the strain tensor, eiJ' namely 'l'xy' 'l'yz and 'l'zx have magnitudes which
are equal to half the magnitudes of the corresponding engineering shear
strains tPxy' tPyz and tPzx•

5.1 ELASTIC STRESS-STRAIN RELATIONS

Hooke first proposed a linear relation between stress and strain for a uniaxial
stress state. A generalised Hooke's law applicable to triaxial stress and
strain states for an isotropic material including thermal strain effects can
be defined in the form
ex= [ {ux- v(uy + uJ}/E] ± (f.T
ey = [ {CTY- v(uz + ux) }/E] ± (f.T
ez = [ {uz- v(ux + uy)}/E] ± (f.T
(5.1)
Yxr = r:xrf2G = {(1 + v)/E}r:xr
Yrz = r:rzf2G = {(1 + v)/E}r:yz
Yzx = r:zxf2G = {(1 + v)/E}r:zx
where E is Young's modulus of elasticity, v is Poisson's ratio, G the modulus
of rigidity, (1. the coefficient of linear thermal expansion or contraction, T
the temperature above or below a convenient datum temperature and
G = E/2(1 + v).
It will be noted that ex in equations (5.1) can be rewritten as follows:
ex= {uxf2G(1 + v)}- (v/E)(uy + uz) ± (f.T
= (uxf2G)- vuxf2G(l + v)- (v/E)(uy + uz) ± (f.T
= (ux/2G)- (v/E)(ux + CTY + uz) ± (f.T
=(ux/2G)-(v/E)J 1 ±(f.T (5.2)
Similar expressions can be obtained for eY and ez. Equations (5.1) can
therefore be defined in tensor notation as
(5.3)
where J 1 = ux + ur + uz and b;J is the Kronecker delta which has the values
bij = 1(i = j)
= O(i -:J= j)
Equations (5.3) can then be transposed to solve for the stresses, CT;J, where
CT;J = 2Ge;J + b;J2G{ (v/E)J 1 +(f.T}
= 2Ge;J + b;J[ {v/(1 + v) }J 1 + 2G(f.T]
= 2G{e;J- b;J( ± (f.T)} + b;J[ {v/(1 + v) }J1 ]
By adding the first three of the equations (5.1) it will be found that
92 ENGINEERING PLASTICITY

I 1 =ex+ ey + ez = {(1 - 2v)(ux +cry+ uz)/E} ± 3tXT


or I 1 = {(1- 2v)/E}J 1 ± 3tXT (5.4)
Therefore
J 1 = {E/(1- 2v)}(J 1 + 3tXT) (5.5)
and {v/(1 + v) }J 1 = {vE/(1 + v)(1- 2v) }(I 1 + 3tXT)
= t/I(I 1 + 3tXT)
where t/1 = vE/(1 + v)(1 - 2v)
Hence u;j = 2G {e;j - c5;j( ± tXT)} + c5;j t/1(1 1 + 3tX T) (5.6)
Since E/(1- 2v) = 3K, where K is the bulk modulus, equation (5.4) can
be rewritten as
I 1 = (J tf3K) ± 3tXT = (umfK) ± 3tXT (5.7)
and em= I tf3 = {(1- 2v)/E}um ±tXT= (um/3K) ±tXT (5.8)
In the discussion it has not been inferred that the values of E, G, v, tX and K
are, in fact, constant throughout the deforming material. Thus, if the tempera-
ture is not uniform and a temperature gradient exists then the elastic para-
meters may have different values corresponding to different points in the
elastically deforming region.
It should be noted that the free thermal strain of a metal is defined as
that part of the strain which is independent of the stress resulting from
contact or surface forces and is only due to the change in temperature. The
following assumptions are also made:

(1) Since the metal is assumed to be isotropic, the free thermal strain
is the same in all directions for the same change in temperature, so that
the free thermal strains ex<tl = ey<t> = ez<t> = ± tXT.
(2) The free thermal strain is directly proportional to the temperature
change but the coefficient of linear thermal expansion or contraction, tX,
may be a function oftemperature.
(3) That the thermal strain effects can be considered in terms of the
principle of superposition.
(4) The shear strains are not affected by changes in temperature.

It is convenient here to distinguish between the strains which produce


a change of shape and those which produce a change in volume of the
elastically deforming metal.
If um is the hydrostatic stress and em is the corresponding volumetric
strain then, neglecting the free thermal strain, the first of the equations
(5.1) may be rewritten in the form

ex= (ux/2G)- (v/E)J 1 from equation (5.2)


STRESS-STRAIN RELATIONS 93
={(ax- am)/2G} + (am/2G)- 3(v/E)am
= {(ax- am)/2G} + { (1 + v)am- 3vam} /2(i(1 + v)
={(ax- am)/2G} + {(1- 2v)/E}am
Therefore
ex= (a~/2G) + {(1 - 2v)/ E}am (5.9)
and the complete elastic stress-strain relations of equations (5.1), neglecting
the thermal strains, can be stated in tensor notation as
(5.10)
Equations (5.8) and (5.3) can be combined by subtracting em given by
equation (5.8) from equation (5.3) thus producing
eii- em= (ai)2G)- bii{(v/E)J 1 +aT}- biJ {(1- 2v)/E}am ±aT]
= (ai)2G)- <5iJ(v/E)3am + {(1- 2v)/E}amJ
= (ai)2G)- bii[(1 + v)am/E]
= (ai)2G)- <5iJam/2GJ
= (aii- am)/2G
Therefore
(5.11)
which relates the deviator strain tensor with the deviator stress tensor,
such that

5.2 ELASTIC STRAIN ENERGY FUNCTIONS


When a metal is subjected to a system of external forces such that it is
elastically deformed then external work is done. If the metal is in equilibrium
and does not possess kinetic energy then all the external work done is stored
as strain energy of deformation. The elastic strain energy per unit volume
for the idealised metal can be stated 2 as

U = (1/2)aiieii (5.13)
Substituting for the appropriate stresses from equation (5.6) gives
+
U = (1/2)[2G{eii- bii( ±aT)}+ biil/1(1 1 3a T)]eii
which can be rearranged to produce
U = Geiieii + (t/1 /2)/i- {(2G + 3t/J)/2}aTI 1 (5.14)
94 ENGINEERING PLASTICITY

which can also be stated as


U= { (2G + 31/f)(I~- 3rxTI 1 )/6} + 2GI~ (5.15)
where I 1 = ex + eY + ez represents the spherical state of strain and is pro-
portional to the volume change and I~ is the second invariant ofthe deviator
strain tensor which represents a pure distortional strain.
Equation (5.15) can therefore be considered as comprising two terms.
The first term is the strain energy per unit volume related to the change
in volume whilst the second term is related to the strain energy per unit
volume involved in shear distortion.
The elastic shear distortion energy per unit volume, designated U d, is
then defined as
(5.16)
However, it has been shown by equations (2.55) and (3.56), respectively, that
J~ = I~ = 0. Also equation (3.57) shows that I~ = (3/2)y&:T and by analogy
J~ = (3/2)-r~CT'
From equation (5.11)

and it follows that


I~= J~j4G 2
Substituting for I~ in equation (5.16) gives
ud = (1/2G)J~ (5.17)
The quadratic invariant, J~, of the deviator stress tensor, u;P is given by
J 2 = - O'xO'y O'yO'z O'zO'xJ + 'rxy
l ( I I+ I I+ 2 + 'ryz
I 2 + 'tzx
1\ 2
= - {(ux- O'm)(O'y- urn)+ (uy- O'm)(O'z- O'm) + (uz- O'm)(O'x- O'm)}
2 2 2
+ 'rxy + 'ryz + 'tzx
where O'm = (ux + O'y + O'z)/3.
Therefore
J~ = - {(O'xO'y + O'yO'z + O'zO'x)- 2um(O'x + O'Y + O'z) + 3u!} + -r;Y + -r;z + -r;x
= (1/3) {(ux + O'Y + O'z) 2 - 3(0'xO'y + O'yO'z + O'zO'x)} + -r;y + -r;z + -r;x
= (1/3) [ (1/2) {(ux- ul + (uy- O'z)2 + (uz- O'x) 2}] + -r;y + -r;z + -r;x
Therefore J~ = (1/6) {(ux- ul + (uy- uzf + (uz- O'x) 2} + -r;Y + -r;z + -r;x
(5.18)
which agrees with equation (4.8)
The representative or equivalent stress, ii, in terms of the components of
the stress tensor, uij, is given by
ii = (1/2)1/2 \r (ux _ uy)2 + (uy _ uz)2 + (u z _ ux)2 + 6(-r2xy + -r2yz + -r2zx) } 112
(5.19)
STRESS-STRAIN RELATIONS 95
Therefore J~ = (a)2/3 (5.20)
and then Ud=(a) 2 /6G (5.21)

5.3 PLASTIC STRESS-STRAIN RELATIONS


During elastic deformation, the stress-strain relations are linear and the
strains are uniquely determined by the stress state without regard to how
the stress state is attained. However, in the plastic range, the relations are
generally nonlinear. The strains are not uniquely determined by the stress
state but depend on the history of how the stress state was reached and
this will now be illustrated.
Consider the initial yield locus to be that shown as the curve AB in figure
5.1. Assume that a material is subjected to a uniaxial stress, ax, initially,
so that it is plastically strained beyond the initial yield condition to some
point C. The subsequent yield locus may then be defmed as the curve CD

}
and the plastic strain state is given by

~
eP=eP=
y Z
-eP/2
X
(5.22)
Y~y = Y:z = Y~x = 0
where the superscript, p, denotes plastic strain.
If the uniaxial stress is now relieved to, say, the point E and the material
is then subjected to a shear stress, •xy• which increases from zero at E to
the point F on the subsequent yield locus, the plastic strain state remains
that defined by equations (5.22).
However, any other stress path may have been chosen to arrive at the
point F from the point C provided that the path did not pass outside the

Figure 5.1. Dependence of plastic strain on stress history


96 ENGINEERING PLASTICITY

subsequent yield locus. For example, the path OCF following the subsequent
yield locus could have been chosen.
Suppose the material to be subjected initially to a shear stress, Txy, to
the point D on the subsequent yield locus and then by an arbitrary stress
path, such as DGF, again stressed to the point F. In this case, the plastic
strain state is given by

(5.23)
and this plastic strain state is seen to be unrelated to that given by equations
(5.22) for the previous case, although the same stress state is defined by
the point F for both paths. The elastic strain states are consequently the
same but the plastic strain states are different.
Because of the dependence of the plastic strains on the stress path it is
usually necessary to consider incremental plastic strains throughout the
stress history and determine the total plastic strain by integration. However,
if a proportional stress path is followed, such that all the stresses increase
in the same ratio, then the plastic strain state is independent of the stress
history and depends only on the final stress state.

5.3.1 The Prandti-Reuss equations


The determination of the plastic stress-strain relations was apparently
originated by Saint-Venant 3 in 1870 with his treatment of plane plastic
strain. He proposed that the principal axes of strain increment and not of the
total strain coincided with the principal axes of stress.
The stress-strain relations for an elastic-perfectly plastic material were
first proposed by Prandtl 4 in 1924 for plane strain deformation and genera-
lised independently by Reuss 5 in 1930. Reuss assumed that the plastic strain
increment is, at any instant of loading, proportional to the instantaneous
stress deviation and the shear stresses such that
de:.'/<r~ = de~/<T~ = de~/<T~ = dy:.'y/Txy = dy~z/Tyz = dy~x/Tzx = dA,
or, more compactly, in tensor notation as
defi = o-;idA, (5.24)
where, o-;i, is the deviator stress tensor and, dJc, is a scalar non-negative
constant of proportionality which is not a material constant and may vary
throughout the stress history. The superscript, p, is used in the equation
to denote the plastic strain increment.
Equations (5.24) indicate that the increments of plastic strain depend
on the current deviator stress state and not on the increment of stress required
to attain this state. It is also implied that the principal axes of stress coincide
with the principal axes of strain increment. It should be appreciated that
equations (5.24) give the ratios of the plastic strain increments in the various
STRESS-STRAIN RELATIONS 97
directions and do not determine the actual magnitudes of the plastic strain
increments. For the magnitudes of the plastic strain increments to be deter-
mined it is necessary to introduce a yield criterion. From equations (5.24)
de~ju~ = d.A.
or de~= u~d.A. = (ux- um)d.A.
= {O"x- (ux + O"Y + uz)/3} d.A.
= (2/3)d.A. {O"x- (1/2)(uy + uz)}
Equations (5.24) can therefore be stated in terms of the components of the
stress tensor, uij• as

de~= (2/3)d.A. {O"x- (1/2)(uy + uz)}

de~= (2/3)d.A. {CTY- (1/2)(uz + ux)}

de~= (2/3)d.A. {O"z- (1/2)(ux + uy)}


(5.25)
dy~Y = d.A.rxy

dy~z = d.A.tyz

dy~x = d.A.rzx

If the value of d.A. were known then the plastic stress-strain relations
could be defined. By subtracting the second from the first of equations (5.25)
produces
de~- de~= (2/3)d.A.{ O"x- (1/2)(uy + uz)- CTY + (1/2)(uz + ux)}
= (2/3)d.A. {(3/2)(ux- uy)}
= d.A.(ux- uy)
Therefore

It then follows that


(de~- de~f +(de~- de~) 2 +(de~- de~) 2 + 6 { (dy~y) 2 + (dy~z) 2 + (dy~x) 2 }
= (d.A.f {(ux- uy) 2 + (uy- uz) 2 + (uz- ux) 2 + 6(r;Y + r;z + r;x)} (5.26)
By analogy with equations (3.50) and (3.51) it will be seen that the left-hand
side of equation (5.26) is related to the representative or equivalent plastic
strain increment, deP, and equals (9/2)(d.<:P) 2 • Similarly, the right-hand side
of equation (5.26) is related to the quadratic invariant, J~, of the deviator
stress tensor, u;p and equals (d.A.) 2 6J~. It has also been shown by equation
(5.20) that J~ = (o') 2 /3. Therefore
(9/2)(dW') 2 = (d.A.) 2 6J~ = 2(d.A.) 2 (a) 2
or d.A. = (3/2)(dePfa) (5.27)
98 ENGINEERING PLASTICITY

and the plastic stress-strain relations of equations (5.25) become


de~= (dBP /ii) {ux- (1/2)(uy + uz)}
de~= (dBP/ii){uY- (1/2)(uz + ux)}
de~ =(dBP/ii){uz- (1/2)(ux + uy)}
dy~y = (3/2)(d8P/if)'t'xy (5.28)
dy~z = (3/2)(dBP/ii)'t'yz
dy~x = (3/2)(dBJ' /ii)'t'zx
For a perfectly plastic material these equations may be stated as
de~= (3/2)(diiP/ii)u;j (5.29)
Equations (5.29) are functions of the quadratic invariant, J~, of the deviator
stress tensor, u;j, and therefore the original Reuss assumptions imply that
the von Mises yield criterion is applicable.
The total strain increment is the sum of the elastic strain increment,
which will now be defined as dee, and the plastic strain increment deP.
Therefore
and by substitution for, de~P and de~, from equations (5.10) and (5.24),
respectively,
deij = (u;j/2G) + oij{ (1 - 2v)/E}um + u;jdA. (5.30)
or deij = (u;/2G) + oij{(l- 2v)/E}um + (3/2)(dBP/ii)u;j (5.31)
by substituting for de~ from equations (5.29).
These equations for an elastic-perfectly plastic material are known as
the Prandtl-Reuss equations and are obviously complex. During many metal
forming processes, the elastic incremental strains are negligible compared
with the plastic incremental strains in the fully developed plastic region.
In these applications it is then possible to neglect the elastic component
of the total incremental strain as being insignificant. When the elastic
component of the total incremental strain is of importance such as in elasto-
plastic deformation and the metal forming problems of springback, which
is associated with elastic recovery, and also residual stresses, it is necessary
to resort to the Prandtl-Reuss equations to produce the required solutions.
If the elastic incremental strain is neglected then the material is considered
as rigid-perfectly plastic, being incapable of elastic deformation and having
an infmite modulus of elasticity. No strain occurs unless the yield stress
is attained and the total strain increment is identical with the plastic strain
increment.
Considering the principal directions, then equations (5.24) can be stated as
(5.32)
STRESS-STRAIN RELATIONS 99

...
~

...."
'-

'-

""
.c
II)

Plastic strain
dE:P increment dE:P
I

Figure 5.2. Mohr circles for stress and plastic strain increment

or de~ = u~ dA., de~ = u~ dA., de~ = u~ dA.


Therefore

(de~- de~)/(u~ - u~) = dA.}


(de~ - de~)/(u~ - u~) = dA. (5.33)
(de~ - den/(u~ - u~) = dA.
The numerators of the left-hand terms of these equations are the diameters
of the three Mohr circles for the plastic strain increment and the denomi-
nators are the diameters of the three Mohr stress circles as shown in figure 5.2.
Equations (5.33) therefore imply that the Mohr circles of stress and plastic
strain increment are similar.

5.3.2 The Levy-von Mises equations


A relation between the ratios of the components of the strain increment
and the ratio of the stresses was originally proposed by Uvy 6 in 1871 and
independently reaffirmed by von Mises 7 in 1913 when the relation became
better known. These equations are usually known as the Levy-Mises equa-
tions and may be stated in the form
dexfu~ = dey/u~ = dezfu~ = dyxy/r:xy = dyyz/r:yz = dyzxfr:zx = dA.
or (5.34)
Because these equations involve the total strain increment and not the
plastic strain increment they are strictly applicable to a rigid-perfectly
plastic material for which the elastic component of the total strain increment
100 ENGINEERING PLASTICITY

is zero. The total strain increment and the plastic strain increment are then
identical. It follows that the Mohr circles of stress and total strain increment
are also identical.
Stated in terms of the components of the stress tensor, uii' the U~vy-Mises
stress-strain relation includes three equations of the type
dex = (2/3)dA. {ux- (1/2)(uy + uz)} =(deja){ u,.- (1/2)(uy + uz)}
and three equations of the type
dyxy = 't"xydA. = (3/2)(d8/0')-rxy (5.35)
or deij = (3/2)(deja=)u;j (5.36)
The Uwy-Mises equations have been considered here as a special form
of the Prandtl-Reuss equations (5.31) although these latter equations were
actually proposed much later. For the solution to many metal forming
problems where unrestricted plastic flow is relevant it is customary to apply
the Levy-Mises equations without incurring significant error.
Equations (5.29) or equations (5.36) are known as the .flow rule associated
with the von Mises yield criterion.

5.4 STRESS-STRAIN-RATE EQUATIONS


Equations similar to equations (5.29) and equations (5.36) can also be derived
involving strain-rates, namely
eij = u;jA. (5.37)
or = (3/2)(lja)u;j
iij (5.38)
Hence ez = (3/2)(e/a)(ux- um) = (eja){ux- (1/2)(uy + uz)}
ey == (3/2)(e/a)(uy- um) = (e/a){uy- (1/2)(uz + ux)}
ez = (3/2)(ejo)(uz- um) = (e/a) {uz- (1/2)(ux + uy)}
(5.39)
Yxy = (3/2)(ija=)'t"xy
Yyz = (3/2)(l/a)-ryz
Yzx = (3/2)(ija=)-rzx
where, following the so-called Newton fluxion notation, a dot placed above
a symbol denotes differentiation with respect to time, t, so that t= de/dt,
t = de/dt, y= dy/dt, etc.
5.5 PLASTIC WORK AND STRAIN-HARDENING HYPOTHESES
When an engineering metal is formed at ambient temperature it usually
strain-hardens, that is, its resistance to further deformation increases and
the stress required to maintain the deformation therefore increases. During
STRESS-STRAIN RELATIONS 101
the deformation external work is done and in plasticity theory an important
concept is that of plastic work.
The work done per unit volume on an element during deformation is
d W = criideii
= crii(de~i +de~)
=dW,+d~ (5.40)
The work done per unit volume, dW, = criide~i' is that required to produce
the elastic deformation and is recoverable. However, the energy dissipated
per unit volume in producing the plastic deformation, d ~ = criide~ is not
recoverable because plastic deformation is an irreversible process. It can
also be written as
d ~ = cr;idefi = cr'1 de~ + a~ dei + a~ de~ (5.41)
and in terms of the principal stresses becomes
d~ = cr 1 de~+ cr 2 dei + cr 3 de~ (5.42)
so that the total plastic work per unit volume during a finite deformation
is given by

f
~ = criide~ (5.43)

where the integration is performed over the actual strain path from the
initial state of the metal.

cr.'3 p
2G.d€ 3

----------fJ= O

Figure 5.3. Deviator stress and plastic strain increment vectors represented on
the n-plane
102 ENGINEERING PLASTICITY

Consider a plot in the n-plane where the axes are the directions of the
principal deviator stresses u~ , u~ and u~. This is possible because, for a
point on the n-plane, u~ = u 1 , u~ = u 2 and u~ = u 3 since the equation for
the n-plane is u 1 + u 2 + u 3 = 0 and urn= 0. In addition, since for plastic
strain increments, de~ + de~ + de~ = 0, the components of the plastic strain
increment can also be represented by vectors on the n-plane provided these
are multiplied by, for example, the constant 2G in order that they have the
dimensions of stress. The deviator stress vector, OP, and the plastic strain
increment vector, say, QR, can then be represented in the same diagram
as shown in figure 5.3.
From equation (5.41) the plastic work increment, d»;,, is given as the
scalar product ofthe two vectors OP and QR. Hence
d»;, = (OP.QR)/2G
= {IOPIIQRI cos(O-y)}/2G
where IOP I= {(u~) 2 + (u~) 2 + (u~)2 P' 2 = (2/3) 1' 2 a
and IQRI = 2G{(de~f + (de~) 2 + (de~) 2 } 1 ' 2 = 2G(3/2) 1' 2 deP
The plastic work increment is therefore given by
d»;, = udiP cos(B- y) (5.44)
If it is now assumed that the principal axes of plastic strain increment
coincide with the axes of principal stress then y = 0 and equation (5.44)
reduces to
d»;, = adeP (5.45)
which is valid for the Prandtl-Reuss stress-strain relations. It then follows
that the equations (5.29) can be restated in the form
de~= (3/2)(d»;,/u 2)u;j (5.46)
It is necessary to determine the degree of strain-hardening or work-
hardening which occurs in a given metal due to plastic deformation. Two
hypotheses have been proposed for this purpose. One of these hypotheses
proposed by Hill assumes that the degree of hardening depends only on
the total plastic work done and is independent of the strain path 8 • This
implies that the resistance to further deformation depends only on the
amount of plastic work which has been done on the metal from the time
that it was in its initially annealed condition and is assessed by means of
the yield criterion.
As shown previously, the yield criterion for an isotropic, strain-hardening
metal can be written as cf>(u;j) = K, where the magnitude of K changes
as the metal strain-hardens but cf>(u;j) remains independent of direction if
the metal is isotropic. According to the hypothesis, K is a function of the
total plastic work done per unit volume, so that
(5.47)
STRESS-STRAIN RELATIONS 103
In chapter 4, it was shown that the von Mises yield criterion predicts
yielding for engineering metals with sufficient accuracy regardless of the
degree of prestrain. The final yield locus according to this criterion is indepen-
dent of the strain path and also the hydrostatic component stress.
In terms of the principal stresses, this isotropic, strain-hardening yield
criterion is
(0"1- 112) 2 + (u2- u3) + (u3- 111)2 = 6k 2
where k is the current yield shear stress depending upon the amount of
prestrain. However, it is more convenient to use the representative or
equivalent stress, ii, to denote the yield function, that is
ii = [ (1/2){ (u 1 - u2)2 + (u 2 - u 3f + (u3 - u 1)2} ]112 = [ (3/2){ (u~)2 +
(u~)2 + (u~)2} ]112 = f(»i,) (5.48)
Equation (5.48) is only valid for isotropic, strain-hardening where the
Bauschinger effect is absent. It is also implied that no plastic work is done
by the hydrostatic components of stress. No change in volume of the metal
occurs even though its shape is changed during plastic deformation. Its
density remains constant and the metal is therefore considered to be in-
compressible.
The strain-hardening hypothesis can then be written in the form

(5.49)

and it follows that ii is a function only of f diP where the integration is


performed over the strain path. Hence

jj f
= H diP = HiP (5.50)

This equation is usually more convenient to use than equation (5.47)


although these two equations may not produce the same result when applied
to an engineering metal which is anisotropic and the Bauschinger effect
may be present. Equation (5.50) is a statement of the second strain-hardening
hypothesis which assumes the equivalent plastic strain, iP as the assessment
of strain-hardening. Equations (5.47) and (5.50) were distinguished by Ford 9
and referred to as the work-hardening and strain-hardening hypotheses,
respectively.

5.6 EXPERIMENTAL VERIFICATION OF THE


PRANDTL-REUSSEQUATIONS
The frrst experiments to verify the plastic stress-strain relations were
performed by Lode in 1926 (reference 9 of chapter 4). In addition to the
stress parameter, J.4 defined by equation (4.37), Lode also introduced the
plastic strain parameter, v, defmed by
104 ENGINEERING PLASTICITY

v = (2de~ -de~ - denf(de~ -den


={de~- (1/2)(de~ +de~)}/{ (1/2)(de~- de~)} (5.51)
If equations (5.32) are correct then J.l should be equal to v. From equations
(5.32)
v = (2de~- de~- de~)!(de~- den
= (2u~ - u~ - u~)/(u~ - u~)
= (2u2- 0'3- u1)/(u3- 0'1) (5.52)
Thus V=J.l
To prevent confusion, it should be appreciated that the Lode plastic strain
parameter, v, is not related to Poisson's ratio.
If the Prandtl-Reuss equations are valid, then equation (5.52) should
be satisfied over the whole range of experiments. The experiments performed
to evaluate Lode's parameters have been for plane stress conditions. As
explained in chapter 4, Lode stressed thin-walled tubes of steel, copper and
nickel in combined tension and internal pressure. The ratio between axial
stress and circumferential stress was maintained approximately constant
throughout each test and the results are represented diagrammatically in
figure 5.4. The relation appears to be approximately satisfied but the deviation
from the theoretical prediction cannot be accounted for by experimental
error. Although part of the deviation can be attributed to anisotropy of
the test materials it is probable that the Prandtl-Reuss equations are in
some error. The deviation was confirmed by Taylor and Quinney (reference 8
of chapter 4) who stressed thin-walled tubes of aluminium, copper and
mild steel in combined tension and torsion. These investigators were extre-
mely careful to limit the degree of anisotropy in the tubes. However, in

Figure 5.4. Diagrammatic representation of the Lode parameters


STRESS-STRAIN RELATIONS 105
these tests, the axial force was maintained constant whilst the couple was
increased. Consequently, the stress ratios were not constant.
It was suggested by Pugh 10 that it is impossible to be entirely certain
that a thin-walled tube is isotropic. With a view to overcoming this difficulty,
Hill (reference 14 of chapter 4) proposed that notched strip specimens
should be used instead of thin-walled tubes to verify plastic stress-strain
relations because the degree of anisotropy can be better controlled. Hundy
and Green 11 obtained very good correlation between their experimental
results, using this method, and the Uvy-Mises stress-strain relations.
Other experiments by, for example, Hohenemser 12 in 1931 and Morrison
and Shepherd 13 in 1950 using thin-walled tubes subjected to various combi-
nations of axial tension and torsion have demonstrated the approximate
validity of the Prandtl-Reuss equations. Thus, for practical purposes, the
Prandtl-Reuss and Levy-Mises stress-strain relations appear to be suffi-
ciently accurate.

5.7 DERIVATION OF THE GENERALISED PLASTIC


STRESS-STRAIN RELATIONS
In sections 5.3.1 and 5.3.2, the Prandtl-Reuss and Levy-Mises stress-strain
relations, respectively, were considered and the equations originated from
the assumption that the principal axes of strain increment coincided with
the principal axes of stress. It was also shown that these relations imply
the von Mises yield criterion to be applicable.
However, it is possible to derive equations for determining the plastic
stress-plastic strain increment relations for any yield criterion employing
a unified approach due to Drucker 14 •15 . When a strain-hardening material
was previously considered it was defined in terms of its resistance to further
deformation increasing as the process of deformation proceeds. A more
precise definition of strain-hardening is required.

5.7.1 Drucker's postulate


Consider the true stress-natural strain characteristic curves shown in
figure 5.5. The material characterised by figure 5.5(a) is strain-hardening
so that an additional stress increment da > 0 produces an additional strain
increment ds > 0. The work done is then given by da. ds > 0. The additional
stress increment does positive work which is represented by the shaded
area beneath the curve. A material having this characteristic is referred to
as being stable.
Figure 5.5(b) suggests that following a maximum stress being attained,
the stress then decreases with increasing strain. For this part of the curve,
da < 0 and da.ds < 0. Consequently, the work done is negative and this
material is unstable.
For the stress increment shown in figure 5.5(c), da > 0, but the strain
106 ENGINEERING PLASTICITY

dO<Ob
C1 dO'>O C1
Jh
r-~
I
I
I I I
I I I
I I I
E: E:
0 dE:>O 0 dE:>O E:

(a) (b) (c)

Figure 5.5. True stress-natural strain (a, e) characteristic curves showing positive
and negative increments of work

decreases with increasing stress so that dO". dB < 0 and again the work done
is negative. This hypothetical case violates the law of conservation of energy.
Only the first case as illustrated by figure 5.5(a) is normally appropriate
to the deformation of engineering metals.
An element of a strain-hardening material is subjected to an initial stress
state, O"ij, and then, by some external 2.gency, the stress state is changed
to O"ii, which is on the yield locus, by an arbitrary loading path inside the
yield locus so that only elastic deformation is involved. Now suppose that
an additional infinitesimal stress, dO"ii, is applied and then relieved. It is
assumed that these changes occur very slowly so that the process can be
regarded as isothermal. According to Drucker's postulate:

:::l
___ L_ --.. . . . .
Subu:quent yield locus

c.,---
............

Initial yield locus~


Figure 5.6. Cycle of loading and unloading
STRESS-STRAIN RELATIONS 107
(1) During the process ofloading, the additional stresses produce positive
work.
(2) For a cycle of application and relief of the additional stresses, the
net work done is positive if plastic deformation occurs and for a strain-
hardening material the work done will equal zero when the deformation
is purely elastic.

Let the initial yield locus be designated :E, as shown in figure 5.6, and the
loading path be A -+ B -+ C. The initial stress state, uij, then corresponds
to the point A which is either inside the yield locus, as shown in figure 5.6,
or on the surface of :E. The point B representing the stress state, O";j, is on the
surface of :E. An infinitesimal stress increment, du;j, is then applied which
is indicated by B -+ C and produces a corresponding strain increment, de;j,
where part of de;j is elastic and part may be plastic, that is, de;j = de~j +de~.
The subsequent yield locus is designated :E'.
Now suppose that the additional stress increment, du;j, is relieved so
that the material unloads and returns to the point A by an arbitrary path
C -+ A. In this event, the elastic strain increment, de~j, is recovered but the
plastic strain increment, de~, is irreversible. It then follows from the second
implication given above for strain-hardening that

f(u;j- u:J)de;j > 0 (5.53)

For the closed loading and unloading path A -+ B -+ C -+ A, the work


done by the additional stresses in producing the elastic strain increment,
de~j, is zero. The second implication then gives

f(u;j - u~)det > 0 (5.54)

Since plastic deformation occurs only on an infmitesimal segment B -+ C


(5.55)
Inequality (5.55) is sometimes known as the local maximum principle
and for a strain-hardening material can only be equal to zero if plastic
deformation does not occur.
For the loading process B-+ C, the first implication leads to
du;jde;j > 0 (5.56)
and for the cycle of loading and unloading B -+ C -+ B, the second impli-
cation leads to
(5.57)
since the work done in producing elastic strains is zero for a closed path.
Inequalities (5.56) and (5.57) represent a mathematical definition of
108 ENGINEERING PLASTICITY

Figure 5.7. Rectangular coordinates ofthe stress increment vector


strain-hardening, and inequality (5.57) is sometimes referred to as the
uniqueness condition.

5.7.2 Components of the stress increment vector


Let the existing stress state for an element be <Y;j· Now assume that an
additional stress increment, d<Y;j, is then applied which produces plastic
deformation. As indicated in figure 5. 7, the stress increment vector, d<Y;j,
can be resolved into two rectangular components, that is, d<J:Y tangential
to the yield locus which corresponds to neutral loading and does not produce
plastic flow and the other, d<Y!jl, which is normal to the yield locus and
produces a plastic strain increment, delj. The plastic strain increment vector,
delj, is, therefore, normal to the yield locus.

5.7.3 Convexity of yield locus


Inequality (5.57) is a scalar product of the stress increment vector, d<Yij•
represented by AB in figure 5.8(a) and the plastic strain increment vector,
delj, which is positive.
d<Jijdelj ~ 0
d<Yij II delJ 1 cos
1 e~ o
or - n/2 ~ e ~ + n/2 (5.58)
The vectors d<Y;j and delj must therefore make an acute angle with each
other. Also, since the magnitude of <Y;j - <Yij can always be made greater
than the magnitude of d<Y;j, and because
(<Y;j - <Yij)delj ~ 0
1()ij - ():; lldetJ 1 cos 1/1 ~ o
or - n/2 ~ 1/1 ~ + n/2 (5.59)
STRESS-STRAIN RELATIONS 109
Convex yield locus

(a)

Obtuse angle

(b)

Figure 5.8. (a) Stress and plastic strain increment vectors for a convex yield locus
where only acute angles (} and 1/J are possible; (b) a yield locus which is not convex
showing that obtuse angles 1/J are possible

which means that the vector aii - ai'j makes an acute angle with the vector
deE for all values of ai'j. This implies that all points representing ai~ must
be located on one side of a plane which is perpendicular to the vector deE,
and because this vector is normal to the yield locus it follows that this
plane is tangential to the yield locus as shown in figure 5.8(a). No vector
aii - ai'j can pass outside the locus and intersect it twice. The yield locus
must therefore be convex. If, however, the yield locus were not convex, as
is the situation indicated in figure 5.8(b), there can exist some points, aii
and ai'j, such that the vector aii - ai'j makes an obtuse angle, t/J, with the
vector defi, in which case the condition (5.59) is violated.

5. 7.4 Generalised form of the plastic stress-strain relations


In deriving the generalised plastic stress-strain relations, the definition
for strain hardening expressed by inequalities (5.56) and (5.57) is implied
and two assumptions are made:
110 ENGINEERING PLASTICITY

(1) At each stage of plastic deformation there exists a loading function


f(uij) so that further plastic deformation occurs only whenf(uij) > K where
both the loading function, f, and the strain-hardening function, K, may
depend on the existing stress state and on the strain history.
(2) That a linear relationship exists between the plastic strain increment
and the additional stress increment, that is,

(5.60)
where Cijkl may be functions of stress, strain and loading history but are
assumed to be independent of the stress increment duk 1•
If plastic deformation occurs, then the first assumption implies that
df(uij) > 0
or (5.61)
Let the existing stress state be uw and an additional stress increment,
dukP be applied so that plastic deformation occurs. As discussed in section
5.7.2, the stress increment vector, duw can be resolved into the two rectan-
gular components, du~V, and du~~>, which are tangential and normal to the
yield locus, respectively. The component du~~l produces a plastic strain
increment deE. From inequality (5.61)
(aj;auk 1)duk 1 = (af/auk 1 ){du~? + du~{} > 0 (5.62)
However, the tangential component stress increment, du~f, does not produce
plastic deformation. Therefore
(af/auk 1 )du~/ = o (5.63)
The normal component stress increment du~~l is proportional to the gradient
off Therefore
(5.64)
where s is a scalar > 0 and then from inequality (5.62) and equations (5.63)
and (5.64) it follows that
(af/auk1)duk 1 = (af/auk 1 )du~~> = (af/a<Jk 1)s(afla<Jk1) > o (5.65)
Hence s = {(af/a<Jk 1)j(ajja<Jmn)(afla<Jmn)}d<Jk1 (5.66)
Comparing equation (5.66) with equation (5.60) and since the tangential
component stress increment d<J~V does not produce plastic deformation it
can be seen that the components of the plastic strain increment, deE, are
proportional to the scalar s. That is
(5.67)
and substituting for s, produces
deE= gi/af/a<Jk 1)d<Jk 1 (5.68)
STRESS-STRAIN RELATIONS 111

where 9;j depends on the stress, strain and history of loading. Inequality
(5.57) can be restated as
da;jde~ = {daW + dalj>}defj;:::: 0 (5.69)
but dal~l produces no plastic deformation so that the additional stress
increment daij = Cdal~l + daljl, for any value of C, will produce the same
plastic strain increment de~. The strain-hardening condition can therefore
be stated as
{Cda!')
IJ
+ da!':'l}de.~;::::
IJ IJ
0 (5.70)

but daljlde~ = 0 and consequently


daW .de~= dal~l9;/8fjoak 1 )dak 1 = 0
However (ofjoak 1)dak1 > 0. Therefore dal~lgij = 0 and it can then be seen
that
(5.71)
where G is a scalar which may depend on stress, strain and the history of
loading.
If the substitution is now made for 9;j given by equation (5.71) into equation
(5.68) then
de~= G(of/8a;)(8f/8ak 1)dak1 (5.72)

or de~= G(of/oa;j)df (5.73)


Consider an isotropic, strain-hardening material for which the von Mises
yield criterion is applicable. Let the von Mises yield function in terms of
the components of the stress tensor, aij• be given by
f = J~ = (1/6){ (ax- ay) 2 + (ay- az) 2 + (az- ax) 2 } + (-r;Y + r~z + r;x) = k 2
Hence ofjoax = (2/3) {ax- (1/2)(ay + az)}
and ofjorxy = 2rxy
Hence de~= G(ofjoax)df
= G(2/3) {ax- (1/2)(ay + az) }df
= (2/3)d.A.{ax- (1/2)(ay + az)}
and dy~Y = (1/2)d¢~Y = (1/2)G{8f/8rxy)df
= dkrxy
where d). = Gdf The other components of the plastic strain increment,
de~, can be obtained in a similar manner.
If the yield function is expressed in terms of the principal stresses
f = (1/6){ (a 1 - a 2)2 + (a 2 - a 3 ) 2 + (a 3 - a 1f}
and ofjoal =(2/3){a1 -(1/2)(a2 +a3)}
112 ENGINEERING PLASTICITY

Two other equations of the same type can be obtained by determining the
partial derivatives, ofjoa 2, and ofjoa 3. Then three equations of the type
dt:l = (2/3)d-i{a 1 - (1/2)(ax + ay)}
can be produced which give the components of plastic strain increment,
dt:fj, in the principal directions. The Prandtl-Reuss equations can thus
be identified if d-i = Gdf and it will be appreciated that the Prandtl-Reuss
equations or flow rule are associated with the von Mises yield criterion.
In equation (5.27), the non-negative parameter, d-i, has been shown to
be equal to (3/2)(d8P /a). Therefore
d-i = Gdf = (3/2)(dsP /a)
or G = (3/2)(dsP /a)/df
However, from equation (5.20),
f=J~=(a) 2 /3
Hence df/da = (2/3)a
or df = (2/3)ada
and G = (9/4){ dsP /(a) 2da} (5.74)
Now assume the Tresca yield criterion to be applicable and let the maxi-
mum and minimum principal stresses be known where a 1 ~ a 2 ~ a 3 . In this
case,
f=(a 1 -cr 3 )/2=k
ofjoal = 1/2, ofjoa2 = 0 and ofjoa3 =- 1/2
Hence dc:l = (1/2)dA
dc:2 = 0
dt: 3 = - (1/2)d-i
It will be noted that the plastic stress-strain relations or flow rule associated
with the Tresca yield criterion is different to that for the von Mises yield
criterion. It can now be emphasised that each yield criterion is associated
with a different flow rule.

5.7.5 The plastic potential


The generalised flow rule of equations (5.73) can also be obtained using the
concept of a plastic potential which was first introduced by Melan 16 • This
concept is based on the hypothesis that there exists a plastic potential
which is a scalar function of stress, say g(a;), from which the ratio of the
components of the plastic strain increment, dt:fj, can be obtained by partially
differentiating g( a;) with respect to the stresses a;i. Thus
STRESS-STRAIN RELATIONS 113
(5. 75)
where d/3 must be a non-negative constant such that a negative strain is not
associated with a positive stress.
If equations (5.75) are compared with equations (5.73) it would appear
that the plastic potential is related to the yield function. It was, in fact,
proved by Bland 17 that they are the same function so that gin equations (5. 75)
can be replaced by fand then equations (5.75) and (5.73) are the same.
Substituting the von Mises yield functionffor gin equations (5.75) leads
to the Prandtl-Reuss equations and the Levy-Mises equations when the
elastic components of the strain increment are negligible compared with
the plastic components of the strain increment. In terms of the principal
stresses
g(ai) = f(ai) = J~ = (l/6){(a 1 - a 2 ) 2 + (a 2 - a 3)2 + (a 3 - a 1f}
ajjaa 1 = (2/3){a 1 - (l/2)(a 2 + a3)}
= a1- am where am= (a1 + az + a3)/3
=a~

Hence dei = (ag;aaij)d/3


= (af I a a 1)df3
= a'1 d/3

or del/ a'1 = . . . . . = d/3 (5.76)


The plastic potential g(ai) must necessarily be a symmetric function of
the three stress invariants and is independent of the system of coordinates
chosen. For a stress state (a 1, a 2 , a 3) on the yield surface in the Haigh-
Westergaard stress space, the direction of the plastic strain increment vector
is the same as the direction of the normal to the g(ai) surface at that point.
This arises because the ratios of the direction cosines of this normal to the
g( aii) surface at the point (a 1, a 2 , a 3) are given by
ag;aa 1 : ag;aa 2 : ag;aa 3
Hence (5.77)
Also, if the material is assumed to be incompressible during plastic
deformation, then
dei +de~+ de~= (ag;aa 1 ) + (ag;aa 2 ) + (ag;aa 3 ) = o (5.78)
If the Tresca yield criterion is assumed to be applicable and a 1 ~ a 2 ~ a3
then g(ai) =(a 1 -a 3)/2 in which case
dei :de~ :de~=a{(a 1 -a 3)/2}jaa 1 :a{(a 1 -a 3)/2}jaa 2 :a{(a 1 -a 3)/2}jaa 3
= 1/2 :0 : - (1/2)

or 1 :0:-1 (5.79)
114 ENGINEERING PLASTICITY

which indicates that plastic deformation occurs only in the plane of u 1 and
u 3 and that the plastic strain increments are equal in magnitude but of
opposite kind.

5.7.6 Flow rule for an anisotropic material


As for an isotropic material, the hypothesis is made that there exists a
plastic potential g(uij) = f(uij) so that the incremental strains may be derived
by partially differentiatingf(uij) with respect to uij. Restating equation (4.42)
2f(uij) = F(uy- uz) 2 + G(uz- ux) 2 + H(ux- uy) 2 + 2L'r;z + 2M-r;x + 2N-r;Y
Therefore of/oux = G(ux- uz) + H(ux- uy). Substituting for of/oux into
equation (5.73) where Gdf = dA. produces
de~= dA.{ G(ux- uz) + H(ux- uy)}
The other components of the plastic strain increment, de~, can be deduced
in a similar manner and can be stated as
de~= dA.{ G(ux- uz) + H(ux- uy)}

de~= dA. {F(uy- uz) + H(uy- ux)}


de~= dA.{ G(uz- ux) + F(uz- uy)} (5.80)
dy~z = dA.L-ryz
dy~x = dA.M-rzx
dy~y = dA.N-rxy
When the elastic strain increments are negligible compared with the
plastic strain increments then equations (5.80) are the total strain increments
for a rigid-perfectly plastic material and are analogous to the Uvy-Mises
relations for an isotropic material. The superscript, p, can then be deleted.
By adding the first three of equations (5.80) it will be found that de~ +de~ +
de~= 0 showing that the equations satisfy the condition of incompressibility.
During plastic deformation of a material the state of anisotropy changes.
Nevertheless it will be assumed that the change in anisotropy is negligible
compared with the initial state of anisotropy. In this case it must be expected
that the closest correlation with theoretical prediction is to be obtained for
a material which is significantly anisotropic before plastic deformation
occurs.
Assuming the state of anisotropy to remain constant, then the yield stresses
must increase proportionally as the material strain-hardens and it then
follows that the anisotropic parameters must decrease in proportion. The
ratios of these parameters will therefore remain constant and it is these
ratios which are determined experimentally, not the individual values of
the parameters. It is known that the strain ratios remain sensibly constant
for some materials and not for others.
STRESS-STRAIN RELATIONS 115
The application of the simplified theory is therefore not expected to
describe the anisotropic behaviour of all materials and the need exists to
determine clearly the limitations in the application to engineering metals.
It was suggested by Hill 18 that the equivalent stress, a, for an anisotropic
material be defined as
a= (3/2) 112 [ {F(cry- 0" 2 + G(crz- crx) 2 + H(crx- cr/ + 2L-r;z +
2)

2M-r:~ + 2N-c;y}/(F + G + H)]l 12 (5.81)


where only ratios of the anisotropic parameters are considered. If anisotropy
is negligible and loading is along the principal axes of anisotropy then it
will be found that equation (5.81) reduces to equation (5.48).
Hill also assumed, by analogy with isotropic theory, that the equivalent
stress, a, is a function of the plastic work. The increment of plastic work per
unit volume for a rigid-perfectly plastic material is
(5.82)
and using equation (4.42) and Euler's theorem on homogeneous functions
(see, for example, the textbook by Sokolnikoff and SokolnikofT19
dW=2fdP=dl (5.83)
From equations (5.80) the following expressions can be obtained for a rigid-
perfectly plastic material
Fdex- Gdey = (FG + GH + HF)(crx- cry)dl}
Gdey- Hdez = (FG + GH + HF)(cry- crz)dl (5.84)
Hdez- Fdex = (FG + GH + HF)(crz- crx)dl
Since dW= ade=dl
the equivalent strain increment, ds, can then be defined as
de= dlja = (3/2) 1 ' 2 (F + G + H) 1' 2 [F {(Gdey- Hde )/(FG + GH + HF) }2
2

+ ... + (2dy;z/L) + ... ]1 12 (5.85)

5.8 THE PRINCIPLE OF MAXIMUM WORK DISSIPATION


If a rigid-perfectly plastic material is stressed such that the principal stress
state is (u 1 , u 2 , u 3 ) then the incremental plastic work done per unit volume is
(5.86)
where deF are the principal plastic strain increments.
The incremental plastic work done per unit volume is also given by
the scalar product of the deviator stress vector, OP, and the plastic strain
increment vector, PQ, shown in figure 5.9. It should be remembered that
the plastic strain increment vector has a direction normal to the yield locus
at the point P. In addition, only the deviator stress components produce
116 ENGINEERING PLASTICITY
u3 Yield locus

Oj n-planc

Figure 5.9. The increment of plastic work done per unit volume is the scalar
product of the deviator stress vector, OP, and the plastic strain increment vector PQ

plastic work and the hydrostatic stress components do no work. The incre-
mental plastic work done can therefore be considered with reference to
the n-plane as illustrated in figure 5.9. Hence
d W = cr~ de~ + cr~ de~ + cr~ de~ = cr; def (5.87)
Now consider another stress state (cr!, cr!, crt) which produces the same
plastic strain increments, de~, satisfies the yield criterion and is represented
by the point P* on the 7t-plane and which is also on the yield locus.
The incremental plastic work done per unit volume for this case is given by
d W* = cr!' de~ + cr!' de~ + cr!' de~ = cr(' def (5.88)
The difference in the plastic work increments per unit volume is
dW- dW* = OP·PQ- OP*·PQ
= (cr~- cr~')de!'
1 1 1
(5.89)
Thus, for an element of volume, d V, subjected to the more general stress
state, crij, which produces plastic strain increments, de~, and also subjected
to the stress state, cr~, which produces the same plastic strain increments,
the difference in the plastic work increments is
(cr;j- cr~')de~.dV (5.90)
Consider a rigid-perfectly plastic body of volume V to be plastically
deformed such that the plastic strain increments are de~ at every point. The
actual plastic work increment, d W, is greater than the plastic work increment,
dW*, required to produce the same plastic strain increments, de~, by the
fictitious stress distribution which also satisfies the same yield criterion,
since
STRESS-STRAIN RELATIONS 117

dW- dW* =I v (rr;i- rrG')dc~.dV (5.91)

or I v (rr;i - rrG')dc~. d V ?o 0 (5.92)

which implies that a rigid-perfectly plastic material deforming plastically


will do so in a manner such as to cause a maximum dissipation of energy.
Inequality (5.92) expresses the principle ofmaximum work dissipation following
Bishop and Hill 20 .
Instead of involving plastic strain increments, equation (5.91) and inequa-
lity (5.92) can be restated in terms

of the plastic strain rates, f;P.,
IJ
and rates

of energy dissipation, W and W*, so that

I v
(rr;i- rrG')e~.dV ?o o (5.93)

REFERENCES
1. Hill, R., The Mathematical Theory of Plasticity, ch. 2, p. 33, Oxford
University Press, London (1950)
2. Sokolnikoff, I. S., Mathematical Theory of Elasticity, p. 84, McGraw-
Hill, New York (1956)
3. de Saint-Venant, B., Memoire sur l'etablissement des equations
differentielles des mouvements interieurs operes dans les corps solids
ductiles au dehi des limites ou l'elasticite pourrait les ramener a leur
premier etat, C. R. Acad. Sci. Paris, 70, 473 (1870)
4. Prandtl, L., Spannungsverteilung in plastischen Koerpern, Proc. 1st
Int. Congr. on Appl. Mech., Delft, p. 43 (1924)
5. Reuss, A., Beruecksichtigung der elastischen Formaenderungen in der
Plastizitaetstheorie, Z. angew. Math. Mech., 10,266 (1930)
6. Levy, M., Memoire sur des equations generales des mouvements
interieurs des corps solides ductiles au dela limites ou l'elasticite pourrait
les ramener a leur premier etat, C. R. Acad. Sci. Paris, 70, 1323 (1870)
7. von Mises, R., Mechanik der festen Korper in Plastisch deformablem
Zustand, Nachr. Gess. Wiss. Gottingen, p. 582 (1913)
8. Hill, R., The Mathematical Theory of Plasticity, ch. 2, p. 25, Oxford
University Press, London (1950)
9. Ford, H., Advanced Mechanics of Materials, p. 416, Wiley, New York
(1963)
10. Pugh, H. Ll. D., A note on a testoftheplasticisotropyofme tals,J. Mech.
Phys. Solids, 1, 284 (1953)
11. Hundy, B. B. and Green, A. P., A determination of plastic stress-strain
relations, J. Mech. Phys. Solids, 3, 16 (1954)
12. Hohenemser, K., Fliessversuche an Rohren aus Stahl bei kombinierter
Zug und Toisionsbeanspruchung, Z. angew. Math. Mech., 11, 15 (1931)
118 ENGINEERING PLASTICITY

13. Morrison, J. L. M. and Shepherd, W. M., An experimental investigation


of plastic stress-strain relations, Proc. lnstn, mech. Engrs., 163, 1 (1950)
14. Drucker, D. C., Some implications of work hardening and ideal plasti-
city, Q. Jl. appl. Math., 7, 411 (1952)
15. Drucker, D. C., A more fundamental approach to plastic stress-strain
relations, 1st U.S. Congress of Applied Mechanics, p. 487, ASME,
New York, (1952)
16. Melan, E., Zur Plastizitaet des raeumlichen Kontinuums, Ingr.-Arch.,
9, 116 (1938)
17. Bland, D. R., The associated flow rule of plasticity, J. Mech. Phys.
Solids, 6, 71 (1957)
18. Hill, R., The Mathematical Theory of Plasticity, ch. 12, p. 332, Oxford
University Press, London (1950)
19. Sokolnikoff, I. S. and Sokolnikoff, E. S., Higher Mathematics for
Engineers and Physicists, p. 136, McGraw-Hill, New York (1941)
20. Bishop, J. F. W. and Hill, R., A theory of the plastic distortion of a
polycrystalline aggregate under combined stresses, Phil. Mag., 42,
414 (1951)
6
Phenomenological Nature of
Engineering Metals

Previous chapters have been concerned with establishing the fundamental


principles of plasticity theory which, together with the theory of elasticity,
describe the mechanics of deformation for most engineering metals. Both
ofthese theories are based essentially on experimental studies of the relations
which exist between stress and strain under relatively simple conditions
of loading. They are therefore on a macroscopic scale and, consequently,
of a phenomenological nature. At the present time, these theories are almost
independent of the microscopic structure and behaviour of metals.
This chapter is concerned with the phenomenological nature of real
engineering metals including the effects of hydrostatic pressure, strain-rate,
temperature, reversal of loading and anisotropy. The stress-strain charac-
teristic curve which is a basic requirement for the application of plasticity
theory is described in some detail. It is thus intended to focus the attention
of readers on the complex behaviour of metals and hence highlight the
various limitations which are imposed when applying any theory.

6.1 CHANGES IN DENSITY AND SHAPE


Substances are usually distinguished as solids, liquids or gases, although
from a physical point of view these distinctions may to some extent be
conditional. The classification may be formulated in terms of the effect
of external forces acting on them, namely, by differences in their resistance
to a change of shape. Liquids and gases offer little reilistance to change
of shape whilst an engineering metal such as mild steel has a relatively high
resistance and therefore requires the application of high magnitude force
to produce changes of shape. Since the main purpose of metal forming
processes is to cause a change of shape of the workpiece to produce the
desired product, without metal removal, this property is of great importance.
However, experiments by Bridgman and others have shown that the
volumetric compression of non-porous solids and liquids is an elastic defor-
mation where the relation between the relative volumetric change and the
120 ENGINEERING PLASTICITY

hydrostatic pressure is very nearly linear 1 •2 . This implies that change in


density of a body results from an elastic deformation. During the forging
of steel its volume may be reduced by 0.6 per cent and copper by 1.3 per cent
when subjected to a hydrostatic pressure of about 980 MPa and the metal
caesium undergoes a temporary change in volume of about 30 per cent
when the hydrostatic pressure is 1450 MPa 3 . Further references to density
changes during metal forming at ambient temperature are to be found in
the paper by Shelton4 •
It has been previously shown in section 2.3.11 that any principal stress
state can be regarded as the sum of a hydrostatic stress component which
is equal to the arithmetical mean of the principal stresses and deviator
stress components equal to the differences between the principal stresses
and this hydrostatic stress. Bridgman and, more recently, Crossland 5 have
shown that no plastic flow is caused by the hydrostatic stress. The plastic
flow is caused by the deviator stress system which is so called because the
magnitudes of these stresses are the values by which the actual stresses
deviate from the hydrostatic stress.
The influence of hydrostatic pressure proves to be important for studies
in the movement of rock at great depths in the earth. A number of references
to the compressibility of rocks is given in the textbook by Nadai 6 and he
refers to tests performed by Hughes and McQueen. They suggest that iron
and nickel in the centre of the earth at very high temperature and very high
hydrostatic pressures of about 480.10 3 MPa are elastically compressed
to less than one-half of their normal volume.
For conventional metal forming processes in which large plastic strains
occur and the hydrostatic stresses are moderate it is possible to neglect
changes in density as being insignificant. The workpiece material can then
be considered as incompressible.

6.2 QUASI-STATIC UNIAXIAL TENSILE TEST


6.2.1 Axial force-extension diagram
The most common mechanical test is the uniaxial tensile test to determine
the stress-strain relation of a material. A specimen of the material to be
tested is machined into a shape suitable for gripping in the jaws of a testing
machine and then subjected to an increasing uniaxial tensile force until
fracture occurs. Although the test appears to be simple it is, in fact, very
complex and must be carefully controlled if the results obtained are to be
meaningful. The specimen is formed with enlarged ends suitable for gripping
so that local stress effects due to gripping are insignificant and the transition
between the enlarged ends and the reduced central test portion of the
specimen must be gradual. To ensure that these and other requirements are
satisfied British Standards have been formulated, for example, reference 7.
Tensile test specimens are either circular or rectangular in cross-section,
the latter being used mainly for the testing of strip or sheet metals.
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 121
Consider a bar of initially annealed mild steel to be subjected to an
increasing tensile axial force, F, in a testing machine at ambient temperature
and at a strain-rate less than 10- 3 per sec so that the process can be regarded
as quasi-static and isothermal. It is essential that the specimen is correctly
aligned in the testing machine to ensure that it is not subjected to a couple
and hence there is a uniform axial stress distribution in the reduced section
of the specimen which contains the test gauge length. Measurement of the
extensions, x, corresponding to the varying magnitudes of axial force may
be determined by means of an extensometer when the extensions are small
and with a pair of dividers and rule for subsequent extensions. However,
modern testing machines are equipped with autographic recorders which
enable an automatic graphical plot of axial force, F, versus extension, x,
to be produced during the test. A typical axial force-extension diagram
for initially annealed mild steel is given in figure 6.1.
The portion of the curve OA indicates that the relation between the
axial force, F, and the extension, x, is essentially linear and the point A
defines the limit of proportionality. During this extension the material is
thus subjected to purely elastic deformation. With further extension this
relation becomes nonlinear although the material remains elastic, so that
if the axial force is relieved the material unloads and the specimen gauge
length reverts to its original dimension. For most engineering metals there
is little difference between the limit of proportionality, A, and the elastic
limit, B. The point B indicates the end of purely elastic deformation and
the initiation of plastic deformation. It is known as the upper yield point
and the upper yield stress is then defined as the axial force corresponding

I
I F

.....
LL

u
ji
I
.2
c
><
I
I
<

I
I
I
0 J Extension x

Plastic or Elastic or
irrecoverable extension recoverable extension

Figure 6.1. Axial force-extension diagram for annealed mild steel


122 ENGINEERING PLASTICITY

to this point divided by the initial cross-sectional area of the specimen


gauge length. Further extension is accompanied by a reduction in the axial
force and then the extension proceeds at approximately constant axial force
corresponding to CD in the diagram. This situation is indicative of a lower
yield stress which is defined as the axial force corresponding to CD divided
by the initial cross-sectional area of the specimen gauge length. The portion
of the curve CD actually represents an average of a series of unstable jumps
between the upper and lower yield points caused by the propagation of
Liiders bands across the specimen. This phenomenon is discussed in some
detail in section 6.2.4. The upper yield point is very sensitive to small bending
stresses, irregularities in the specimen and also the rate of extension. Very
little plastic flow occurs at the upper yield point. The lower yield point
is therefore used for plastic flow calculations.
As the extension continues corresponding to the part of the curve DE,
the axial force required also increases indicating the increasing resistance
of the metal to further plastic deformation even though there is a uniform
reduction in cross-sectional area of the specimen gauge length. The material
therefore strain-hardens. The stress required to produce this further plastic
deformation is usually referred to as the flow stress. However, the slope
of the curve from D to E decreases showing that the rate of strain-hardening
is unable to keep pace with the rate at which the cross-sectional area reduces.
Consequently, at the point E, a maximum value of axial force, Fmax' is
attained followed by a local extension in the gauge length to form a waist
or neck. This situation is only indicative of a condition of instability in
uniaxial tension when uniform extension and therefore homogeneous
deformation ceases. Localised extension then occurs and the uniaxial stress
system changes to a triaxial stress system which is time dependent as further
extension proceeds to fracture defined by the point F. The magnitude of
the axial force required to cause further extension beyond the point of
instability is reduced because the cross-sectional area in the neck rapidly
decreases. The material outside the region of the neck is therefore able to
unload elastically.
If the specimen gauge length is extended significantly, say, to that corres-
ponding to point G in figure 6.1 and then completely unloaded it would
be found that the relation between the axial force, F, and the extension,
x, during unloading is substantially linear and appears as the line GH which
is parallel to the elastic deformation line OA. After completely unloading,
the gauge length is permanently increased. The extension HJ is elastically
recovered whereas the extension OH represents the plastic or irrecoverable
extension. The total extension therefore comprises two components, namely,
the recoverable elastic extension and the irrecoverable plastic extension.
Assuming the specimen is now reloaded, then the unloading line HG is
retraced with a minor deviation. During an unloading and reloading phase
there is generally a small energy loss. This is characterised by a narrow
hysteresis loop being formed which is usually neglected. Plastic flow does
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 123
not occur again until the point G is reached, when further plastic extension
is induced. With further loading the axial force-extension curve continues
along GE as though no unloading occurred.
Most engineering metals and alloys do not exhibit the characteristic
initial yielding of annealed mild steel. In such cases, the transition from
purely elastic deformation to elastic-plastic deformation is gradual and
there is no clearly defined yield point. The axial force-extension diagrams
otherwise have generally the same form as that of figure 6.l.lt is then custo-
mary to defme an arbitrary proof stress in place of the yield stress. The
proof stress is given as the stress which is the axial force divided by the
initial cross-sectional area required to produce a permanent extension equal
to a specified percentage of the initial gauge length. However, proof stress
is not used in the application of plasticity theory.
During plastic extension of a tensile test specimen, Farren and Taylor 8
found that only about 85-90 per cent of the heat equivalent of the external
work reappears as heat. The remainder of the energy is retained in the
deformed crystal lattice of the strain-hardened metal.

6.2.2 Stress-strain characteristic curves for a metal in uniaxial tension


A stress-strain characteristic curve may be easily derived if the axial force-
extension diagram for a metal has been determined. Before proceeding
with the discussion it will be helpful if the quantities are defined which
can be determined from the results of a uniaxial tensile test.
The nominal stress, a nom' is defmed as the magnitude of the current axial
force, F, divided by the initial cross-sectional area, A 0 , of the gauge length.
Nominal stress is therefore given by
(6.1)
The particular value of this nominal stress corresponding to the point
of instability when the maximum axial force, F max> is attained is referred
to as the ultimate tensile strength or ultimate stress which will be designated
CTuu. Hence

(6.2)
If the current cross-sectional area of the gauge length is A, then the true
stress, a, is given by
CT=F/A (6.3)
Let the initial gauge length be, 10 , and the current gauge length be I.
Conventional or engineering strain, e, is defined as the change in length per
unit initial gauge length so that the tensile conventional or engineering
strain is given by
e = (l-1 0 )/10 = (l/1 0 ) - 1 (6.4)
or l/10 = 1 + e (6.5)
124 ENGINEERING PLASTICITY

The percentage uniform elongation is given by


(/inst -/0 )/10 .100 = lOOeinst (6.6)
where linst denotes the current gauge length and einst is the engineering
strain corresponding to the point of instability.
The increment of logarithmic or natural strain, dB, introduced by Ludwik,
is based on the current gauge length and is defined as
de= dl/l (6.7)
and the total logarithmic or natural strain, B, when the initial gauge length,
10 , is extended to a current length, l, is then defined as

B= ('de= rl (d/j/) = ln(/j/ 0 ) = ln(1 +e) (6.8)


Jo Jlo
Hence e = exp(e)- 1 (6.9)
For small extensions, the values of e and Bare almost identical but diverge
as the extension increases.
If the small changes in volume which occur during a tensile test are
neglected as being insignificant and the material considered as incompressible
A 0 l 0 = AI = a constant (6.10)
and A= A 0 (l 0 /l) = A 0 /(1 +e) (6.11)
Since a = F /A from equation (6.3)
a= (F/Ao)(l/lo) = anom(1 +e) (6.12)
It will now be appreciated that a nominal stress (J nom -engineering strain
e characteristic curve is therefore an axial force-extension diagram for a
tensile test specimen having unit initial cross-sectional area for the gauge
length and unit initial gauge length. However, the true stress a-natural
strain B curve is more informative for plasticity applications. The natural
strain has a number of advantages. Natural strains are additive but engineer-
ing strains are not. If a ductile metal is tested in compression and in tension,
the true stress-natural strain curves are almost identical, whereas they are
quite different if engineering strain is used.
The slope of these curves in the elastic region defines Young's modulus
of elasticity, E, so that
E =stress/strain= (F /A 0 )/(x/l 0 )
= Fl 0 /A 0 x (6.13)
where x denotes the purely elastic extension.
For most metals at ambient temperature and for quasi-static deformation,
E is of the order 10 5 MPa and the yield stress, Y, is of the order 10 2 MPa,
so that Y IE is of the order 10- 3 . Plastic deformation is therefore initiated
for many metals when the tensile strain exceeds this value. A typical natural
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 125
strain at the point of instability in uniaxial tension for annealed mild steel
is about 0.25. The extension scale in figure 6.1 is thus purposely distorted
for the sake of clarity.
A further consequence of the condition of incompressibility or constant
volume can be expressed, say, in terms of principal natural strains as
(6.14)
Assuming e1 to be the uniaxial tensile strain during a uniaxial tensile
test, then the transverse strains, e2 and e3 , are equal in magnitude but
compressive, or
e2 =e 3 = -ed2 (6.15)
whereas in terms of engineering strains
(6.16)
which reduces to
e1 + e2 + e 3 = 0
only when the engineering strains ep e 2 and e3 are very small such that
their products can be neglected.
In terms of cylindrical coordinates, the condition of incompressibility
can be stated as
(6.17)
where e. is the natural radial strain, e9 is the natural circumferential strain
and ez is the natural axial strain.
If the true stress-engineering strain (cr, e) and nominal stress-engineering
strain (cr80m,e) curves are drawn for the uniaxial tensile test then their
relative positions are as shown in figure 6.2. The ordinate for cr for a given
(CT,e)

.
0
tf (T

. ..
:: b
.. Ill
(Tno•

"'
~~
"
"'cw:"
-~ .
.. Ill

~
·e;
0 ..
Zl-

0 e
Tensile engineering strain e

Figure 6.2. Stress-strain characteristic curves for a metal in uniaxial tension


126 ENGINEERING PLASTICITY

value of e is given by the product of the corresponding ordinate for unom


and (1 + e) so that for small values of e they are approximately equal but
diverge as the value of e increases.
If the true stress-natural strain (u, e) curve is drawn then this curve is
essentially the same as the (u, e) curve for the same material as presented
in figure 6.2, up to and slightly beyond the yield point. Beyond this point
the two curves diverge. The true stress, however, always increases and does
not have a maximum value at the point of instability.

6.2.3 Strain-rate in uniaxial tension and compression


Tensile engineering strain-rate, e, is the time derivative of the tensile engineer-
e
ing strain, e, so that = de/dt. The increment in tensile engineering strain
de= dl/1 0 (6.18)
Therefore
~ = (dl/1 0 )/dt = (dl/dt)/1 0 = v/10 (6.19)
where t denotes time and v the speed at which the gauge length, 10 , increases.
The tensile natural strain-rate, 8, is
8 = de/dt = (dl/dt)/1 = vjl (6.20)
where l is the current gauge length.
In section 6.2.1 it was stated that for the uniaxial tensile test to be consi-
dered as quasi-static, the strain-rate should not exceed, say, w- 3 per

second.
Assuming the gauge length of a standard tensile test specimen to be
50 mm then v/1 0 ~ w- 3 per sec which indicates that the rate of extension
should not exceed about 50 x 10- 3 mm per sec or 0.3 em per min.
Now suppose that a short cylindrical block of metal of initial height,
H 0 , is subjected to uniaxial compression between two rigid, parallel platens.
For convenience, it will be assumed that the lower platen is stationary
whilst the upper platen descends with a speed, v, and the initial height of
the cylindrical block is reduced to a current height, H, in a small increment
of time dt. The increment of engineering strain is
de= -dH/H 0
and the minus sign thus indicates that the increment of strain is compressive.
The engineering strain-rate, e, is
e= dejdt = - (1/H 0 )(dH/dt) = - v/H 0 (6.21)
The corresponding natural strain-rate, e, is
8 = dejdt = - (1/H)(dH/dt) = - v/H (6.22)
If the cylindrical block were subjected to compression, for example,
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 127
beneath a forging hammer then the strain-rate is variable being a maximum
at the instant of impact and decreasing in some manner until the compression
is completed when the strain-rate is zero.
Let the hammer impact the block with an initial speed, u, and assume
that all the available kinetic energy is actually dissipated in producing
homogeneous plastic deformation of the block so that the compression
from an initial height H 0 to a fmal height H 1 is completed in a time T. The
following mean values of strain-rate can then be defined:

(1) Mean compressive engineering strain-rate is J =efT, where e is the


compressive engineering strain. Therefore
J = ef {(H0 - H 1 )f(uf2)} = {(H 0 - H 1 )fH 0 }f{(H0 - H 1 )f(uf2)}
or J = uf2H 0 (6.23)
suggesting that, J, is one-half of the initial compressive engineering strain-
rate.
(2) Mean compressive natural strain-rate is given by l = efT, where
e = 1n (H ofH 1 ). Therefore
l = ln(H 0 fH 1 )f{(H 0 - H 1 )f(uf2)}
or l = (uf2) ln(H 0 fH 1 )f(H0 - H 1) (6.24)

It should be appreciated that equations (6.23) and (6.24) assume the


deformation to be homogeneous, that the hammer speed decreases in direct
proportion to the amount of compression and there is no loss of energy on
impact. It is usually found that the actual mean compressive engineering
strain-rate is greater than one-half of the initial compressive engineering
strain-rate.

6.2.4 Upper yield point for annealed mild steel


Consider a rectangular section uniaxial tensile test specimen of annealed
mild steel strip to be subjected to an axial force and extended to the upper
yield point B indicated in the axial force-extension diagram of figure 6.1.
If a further increment of extension is caused, then across a particular section
of the specimen it will be found that yielding occurs and a so-called Liiders
band appears. This is illustrated diagrammatically in figure 6.3(a). A Liiders
band is a lamellar zone of plastic deformation which is visible as a grey-black
band inclined at a certain angle, (}, to the direction of the uniaxial tensile
stress. Although the material inside the zone is plastic the remainder of the
specimen material is only elastically strained. The plastic flow is therefore
constrained by the eiastic regions which do not permit plastic strain in the
direction of the length of the band. The direction along which the band
propagates is consequently that of zero extension in the plane of the strip.
128 ENGINEERING PLASTICITY

_Plastic
Liiders
band

(a) (b)

Figure 6.3. (a) Propagation of a plastic Liiders band in an annealed mild steel
strip specimen; (b) engineering strains for an element of the strip in a Liiders band

It is believed that this phenomenon is caused by the precipitation of


iron nitride particles from the supersaturated ex solution which present
obstacles to the movement of dislocations, thus retarding the process of
yielding. The separation of iron nitride from the ex solid solution is slow.
It is well known that this precipitation takes place much more readily
in a distorted crystal lattice so that during straining the iron nitride is
readily precipitated. This phenomenon has also been observed in certain
alloys of aluminium and the ex-/3 brasses with small f3 inclusions. However,
the precipitates causing the obstruction to the movement of dislocations
are not iron nitride.
A Liiders band requires a lower magnitude of stress for its propagation
than its formation. Thus a decrease in stress to the lower yield point is
observed to coincide with the propagation of a Liiders band across the
specimen and during the lower yield point extension CD in figure 6.1 other
bands occur. The deformation is not uniform and the axial force-extension
curve does not begin to rise until the bands have spread throughout the
specimen. Liiders bands do not appear if the metal is prestrained unless
it is strain-aged, that is, allowed to rest after prestraining.
It will be recalled from section 6.2.2 that, provided the engineering strains
are very small,
e 1 + e2 + e3 ~o

and if e 1 is the principal engineering strain in the direction of the uniaxial


tensile stress, the engineering strains in the other two principal directions
are then
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 129
For the element of the strip shown in figure 6.3(b), let OP be of unit
length initially which with further extension becomes OP 1 . However, for
zero extension OP = OP 1 . Hence
QP 1 =cos()+ e 1 cos()= (1 + e 1) cos()
and OQ =sin()- (et/2) sin()= {1- (et/2)} sin()
Since0P 1 = OP = 1
(1 + e 1 ) 2 cos 2 () + {1- (et/2) }2 sin 2 () = 1 (6.25)
If the terms in ei are neglected, then equation (6.25) becomes
(1 + 2e 1 ) cos 2 fJ + (1- e 1) sin 2 () = 1
which simplifies to give
cos 2 () = 1/3
or cos()= ± 1/3 112
and () = ± 54°44' (6.26)
The Liiders bands thus form at approximately ± 55° to the axis of the
specimen.
On either side of an individual Liiders band there is a very thin plastic
wave front of thickness, d, which advances slowly at a speed u and as the
band is propagated the increment of engineering strain is de.
The tensile engineering strain-rate, e, is then given by
e= dej(dju) (6.27)
which implies that an increment in engineering strain, de, occurs in a time
t = dju. Thus if u ~ 1 cm/s, de= 10- 2 , and the thickness of the plastic wave
front corresponds to a grain size of about 10- 2 em, then the strain-rate,
e~ 1/s, which is high for a quasi-static, uniaxial tensile test.
The difference in the values of the upper and lower yield point stresses
is thus sensitive to the strain-rate and also the stiffness of the testing machine.

6.2.5 Empirical equations for stress-strain curves


For some problems in plasticity it is useful to represent the stress-strain
relation for a given material by an empirical equation obtained by fitting
the experimental data. In presenting the following equations, the natural
strain, e, is implied but if the plastic strains involved are of the same order
as the elastic strains it is then preferable to use engineering strain, e.
One of the first empirical equations proposed was that by Ludwik 9 which
has the form
a= Y + Aen (6.28)
where Y is the initial yield stress and A and n are material constants.
130 ENGINEERING PLASTICITY

U U= Y+AE:n

(a)

0
n=l
n•l/2

n•O

(b)

(c)

Figure 6.4. Empirical equations for true stress-natural strain curves: (a) a rigid-
strain hardening material a= Y + Ae"; (b) a= Ae"(Y = 0, n = 1, n = 1/2, n = 0); (c) the
stress-strain curve approximated by a bilinear expression a = Ee from zero strain to the
yield stress Y and a = Pe during plastic deformation

When n = 1, the equation represents a material which is rigid up to


the initial yield stress, Y, and then plastic deformation occurs at a constant
rate of strain-hardening, A, such that the material is linear strain-hardening.
Figure 6.4(a) illustrates the case for 0 ~ n ~ 1 and assumes that the elastic
strains are negligible so that the material can be considered as rigid up
to the initial yield stress, Y.
Assuming Y = 0, then the curves are as shown in figure 6.4(b) and
(6.29)
It should be noted that if the experimental data are plotted logarithmically
then equation (6.29) plots as a straight line. The intercept on the ln a axis
gives the value of ln A which determines A, sometimes referred to as the
strength coefficient, and the slope of the straight line gives n, known as the
strain-hardening exponent. When n =I= 1, the value of da /de ate= 0 is infinite
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 131
which indicates that equation (6.29) should not be used when the strains
are small. However, the equation is applicable to the plastic deformation
of a strain-hardening material at ambient temperature provided the plastic
strains are not very large. If the plastic strains are small then the following
equation is preferable
(6.30)
where B and m are material constants and e is the engineering strain.
The true stress-natural strain curve may also be approximated by using
two or more linear expressions between stress and strain. That all engineering
metals are capable of some elastic deformation can be recognised by a = Ee
where E is Young's modulus of elasticity from zero strain up to the initial
yield stress, Y, and during plastic deformation by a = Pe where P may be
considered as the plastic modulus. Hence
a Ri Ee + Pe (6.31)
and this is illustrated by figure 6.4(c).
When the plastic strains are large, an expression which is known to fit
the experimental data for many engineering metals better than those of
equations (6.28) and (6.29) is that proposed by Swift which in its simplified
form is given by
a= C(D + e)q (6.32)
where 0 ~ q ~ 1 and C, D and q are constants for a given material. The use
of this equation may result in greater difficulties with algebraic manipulation.
One empirical equation which gives a good fit to the stress-strain data
for engineering metals, particularly commercially pure aluminium, is that
suggested by Voce 10 and can be written in the form
a= F + G{l- exp(- re)} (6.33)
where F, G and rare constants for a given material.
Because of its complexity, equation (6.33) is not often used in theoretical
analyses.
A form of equation proposed by Prager is
a= Y tanh(Ee/Y) (6.34)
This curve has a tangent modulus of E at zero strain and the initial yield
stress Y is approached asymptotically at a fast rate.
Another equation which is sometimes used is that due to Ramberg and
Osgood 11 and is given by
e =(a/E)+ k(a / E)P (6.35)
where E is Young's modulus of elasticity and k and p are other material
constants.
132 ENGINEERING PLASTICITY

6.2.6 Plastic instability in uniaxial tension


In section 6.2.1, the phenomenon of plastic instability in uniaxial tension
was discussed, where the axial force applied during a uniaxial tensile test
attains a maximum value defined as the point E in the axial force-extension
diagram of figure 6.1. The true stress corresponding to this condition of
instability can be determined in the following manner. With reference to
figure 6.1 and from equation (6.3)
F=aA (6.36)
Therefore
dF ldx = (a.dAidx) + (A.daldx) (6.37)
At the condition of instability in uniaxial tension dF ldx = 0. Hence
dF = adA + Ada = 0 (6.38)
or da I a = - dA I A (6.39)
Also, if the small change in volume which occurs during a tensile test is
neglected, then the material can be assumed incompressible and equation
(6.10) is applicable. Differentiating this equation leads to

Adl +IdA= 0 (6.40)


or dill= - dAIA (6.41)
Combining equations (6.39) and (6.41) gives
dala = - dA/A = dljl (6.42)

but from equation (6.7), dill= de so that equation (6.42) becomes


dala =de
ot dalde =a= al1
If the true stress corresponding to the condition of instability in uniaxial
tension is denoted by ainst then
dalde = ainst = ainstl1 (6.43)
Since a= anom(l +e) from equation (6.12)
drrlde = anom = al(1 +e)
However, the nominal stress, anom• corresponding to the condition of
instability in uniaxial tension is the ultimate stress, auil' defined by equation
(6.2), Hence
(6.44)
where, einst> is the engineering strain at instability.
If a graph is drawn showing the relation between the true stress, a, and
the natural strain, e, then the true stress, ainst, and the natural strain, einst,
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 133

U Pointlof in:~::lity
-

o__-I.,IE:inst
1------_:__ _ _
(a)

u Point of instability

Oinst

~~-----~~0+---+-----~C
Cinst

(b)

Figure 6.5. Stress-strain curves in uniaxial tension showing construction for


determining the point of instability: (a) true stress-natural strain (a, e) curve; (b) true
stress-engineering strain (a, e) curve

corresponding to the point of instability can be determined. This is achieved


by drawing a tangent to the (a, e) curve, as illustrated in figure 6.5(a), such
that the length of the sub-tangent measured along the natural strain axis
is unity, thus satisfying equation (6.43). Figure 6.5(b) shows a similar cons-
truction which enables the true stress, ainst, and engineering strain, eins 1 ,
to be determined from the (a, e) curve where the sub-tangent measured along
the engineering strain axis is 1 + einst to satisfy equation (6.44).
Assuming that the plastic stress-strain characteristic curve for a given
strain-hardening material at ambient temperature is defined approximately
by the empirical equation (6.29) then
dajde = nAen-l = nAel!je
but a = Aen, therefore
(6.45)
134 ENGINEERING PLASTICITY

However, at the condition of instability in uniaxial tension, equation (6.42)


gives
dcr /de = crinst
Therefore
or einst = n (6.46)
which is the strain-hardening exponent. By substituting this value of einst
into equation (6.9) produces
einst = exp(n)- 1 (6.47)
For most engineering metals, n < 0.5 which indicates that the natural
strain at instability in uniaxial tension is relatively low. This phenomenon
therefore presents a problem during metal forming where the stress state is
predominantly tensile as, for example, in drawing of sheet and bar material.
However, a more serious problem is encountered during sheet metal press
forming operations due to thinning of the metal leading to the initiation
and propagation of fracture. The condition of instability in tension thus
represents a limit to forming.
If the empirical equation (6.30) is used to define the stress-strain charac-
teristic curve approximately, then using equation (6.44) together with equa-
tion (6.30) leads to
einst = m/(1 - m) (6.48)
Substituting this value into equation (6.8) then gives
einst = ln(1 + einst) = ln {1/(1- m)} (6.49)

6.3 QUASI-STATIC COMPRESSION TESTS

6.3.1 The axisymmetric compression test


The order of effective strain effected in many metal forming processes is
appreciably greater than unity. During a uniaxial tensile test of a ductile
metal the range of natural strain obtainable, before the condition of instability
is reached, is much less than unity unless the test is performed with a high
superimposed hydrostatic pressure following the Bridgman method 12 • One
alternative is to extrapolate the tensile stress-strain curve to higher values
but this is regarded as unsatisfactory.
There is evidence to suggest that the quasi-static, true stress-natural
strain (cr, e) curve for a given metal in uniaxial compression corresponds
closely with that in uniaxial tension. It is therefore preferable to perform
a slow compression test on a specimen of the material when data are required
corresponding to large strains. However, the uniaxial compression test
has the main disadvantage that frictional resistance is likely to occur at the
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 135
F

Moving t Platen

r- ------------- -,
I
I
I
I
I
I
I
I
I
I

Stationary

Figure 6.6. Homogeneous compression of a solid cylindrical specimen

interfaces between the specimen and the tooling used to effect the
compression.
The simplest form of compression specimen is a solid circular cylinder
having plane end faces which is compressed between plane parallel platens
that have been hardened and tempered and then ground and polished. To
minimise frictional resistance at the interfaces a suitable boundary lubricant
such as graphite in tallow is used at low temperatures. The machining of
shallow concentric grooves in the end faces of the compression specimen
to entrap lubricant aids in reducing the frictional resistance and this method
was apparently first suggested by Loizou and Sims 13 •
Assuming it to be possible to eliminate frictional resistance at the inter-
faces between the end faces of the specimen and the platens, then the com-
pression of the specimen would be homogeneous. In this event the circular
cylinder of initial diameter, D0 , and initial height, H 0 , would simply be
reduced in height at any instant during the compression to a current height,
H, whilst the diameter increases to D, as shown in figure 6.6. The fractional
reduction effected is then given by
R = (H 0 - H)/H 0 = 1- (H/H 0 } (6.50)
or H/H 0 = 1-R
The compressive natural strain is
e = ln(H 0 /H) = ln{1/(1- R)} (6.51)
Let A 0 be the initial cross-sectional area of the cylindrical specimen and
A the current cross-sectional area. Neglecting the small change in volume
of the specimen which occurs during elastic deformation so that the material
136 ENGINEERING PLASTICITY

Dead metal zone (D.M.Z.) F

Stationary I Platen

(a)
t
(b)

Figure 6. 7. (a) Free compression of a circular cylinder of initial height to diameter


(H 0 /D 0 ) ratio greater than unity with dead metal zones present due to frictional resistance
at the specimen-platen interfaces; (b) diagrammatic representation of the zones which
can be identified and barrelling which occurs during the free compression of a circular
cylinder with friction present at the specimen-platen interfaces (H 0 /D 0 Ri 1)

can be considered as incompressible, then


A 0H 0 = AH = a constant (6.52)
Therefore (6.53)
If the uniaxial compressive force exerted on the specimen by the platens
at any instant is F, then the true compressive stress is
(J=F/A (6.54)
How~ver, even with the most efficient lubrication, there is usually some
frictional constraint at the interfaces between the specimen and the platens.
Material adjacent to the platens may thus be prevented from sliding radially
outwards and in the most severe case approximately conical zones in contact
with the platens are created in which there is little or no plastic deformation.
These zones are usually referred to as dead metal zones. As the compression
proceeds and the platens approach each other the remaining metal is caused
to move radially outwards by the dead metal zones which results in barrelling,
as illustrated in figure 6.7. Figure 6.7(a) shows a specimen of initial height
to diameter (H 0 /D 0 ) ratio greater than unity where the dead metal zones
are formed in contact with the platens but where nq interpenetration of
the dead metal zones has occurred. Three zones are shown diagrammatically
in figure 6.7(b) which can be identified during the compression of a circular
cylinder where the initial (H 0 /D 0 ) ratio is about unity. Zone 1 represents
the dead metal zones which have little movement relative to the platens.
Zone 3 is an annular region which moves mainly in a radial direction and
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 137
Zone 2 is the region where the plastic deformation occurs as material moves
from the interfaces with the dead metal zones into Zone 2 and then into
Zone 3. It should be appreciated that as the platens approach each other
material which was originally on the cylindrical surface of the specimen
moving in a radial direction may ultimately appear at the specimen-platen
interfaces. In such a case the material would have been caused to rotate
through a right-angle. It is therefore necessary at the commencement of a
compression test to lubricate the cylindrical surface of the specimen in
addition to the end faces and the platens.
To minimise the effects of frictional resistance so as to closely approximate
to homogeneous deformation, a cylindrical specimen of initial (H 0 /D 0 )
ratio equal to unity may be used which is subjected to incremental compres-
sion. The compression proceeds noting the magnitude of the axial
compressive force and the corresponding height of the specimen until the
height of the specimen is reduced by approximately 10 per cent. The compres-
sion is then interrupted and the specimen and platens cleaned with a suitable
degreasing agent. The specimen is then remachined to reduce the diameter of
the deformed cylinder such that its (H j D) ratio is returned to unity. After
relubricating all of the specimen and the platen surfaces the compression is
continued and the procedure repeated. In this manner large compressive
natural strains can be effected without encountering barrelling. The author

~/00 = I

~/00 = 3/2
HiD0 = 2
-::.:.:::.1-=----
u..

..."u Derived curve


.....0
0
)(
<{
Fractional
0

.....
-~
0 1/2

-
r.0 2/3
00
Extrapolated value of
fractional reduction R
corrnponding to an
0
..
c::
infinite initial H I D
ratio (00 /H0=0)

Figure 6.8. Extrapolation procedure for the uniaxial compression test suggested
by Cook and Larke 14
138 ENGINEERING PLASTICITY

has obtained compressive natural strains in excess of 2 without undue


difficulty when testing various metals in compression in this manner.
Other methods have been employed in an effort to surmount the difficulties
which can result in inhomogeneous deformation of the specimen during
compression. These include, for example, using three cylindrical specimens in
series and obtaining data from only the central one and using specially
shaped conical platens to offset the formation of the dead metal zones. If
the latter method is employed then there is the difficulty of deciding on
the conical profile to be used.
It would appear that the ideal cylindrical compression specimen is one
having such a height that the end effects are insignificant. There is, however,
a limit to the height of the specimen if buckling is to be prevented. Perhaps
the most notable of the various methods is therefore that suggested by
Cook and Larke 14• A number of specimens of different initial (H 0 /D 0 )
ratios are tested with constant conditions of lubrication and the results
extrapolated in a systematic manner to produce the desired stress-strain
curve for a specimen of infinite initial (H 0 /D 0 ) ratio. The procedure is
illustrated in the graph of figure 6.8.

6.3.2 The plane strain compression test


The axisymmetric compression test described in the previous section is
appropriate when it is possible to machine cylindrical specimens from the

Plat~En Normal plat~En pressure

Specimen strip

Figure 6.9. Arrangement for the plane strain compression test


PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 139
material to be investigated. However, the material to be tested is often in
the form of thin sheet. It then becomes difficult to manufacture suitable
cylindrical specimens. In such circumstances, it is most useful to employ a
compression test performed under conditions of plane strain deformation
which was suggested by Nadai 1 S, in 1931, and later developed by Watts
andFord 16 .
The concept of plane strain deformation is of considerable theoretical
importance and will be explained in some detail in chapter 7. For the present
purposes it will be sufficient to state that this mode of deformation implies
that the strain in one principal direction is zero and thus, in this direction,
the dimension remains unchanged. This is approximately true, for example,
during cold metal strip rolling when the change in thickness of the strip
is accompanied by elongation in the direction of rolling with very little lateral
spread.
A strip of metal of initial thickness, t 0 , and width, w, is indented between
narrow platens of breadth, b, which overlap the strip in its width direction,
arranged as shown diagrammatically in figure 6.9. The platens may be
manufactured from hardened and tempered alloy steel ground accurately
square and parallel and mounted rigidly in a sub-press. A standard sub-press
provided with linear ball race sleeves is very suitable if the force is applied
axially via a lubricated ball seating. If specimens of materials such as stainless
steel or tool steel are to be tested then tungsten carbide platens would have
to be used.
Plane strain conditions are approximately achieved because the specimen
material on each side of the platens remains elastic and therefore constrains
the plastically deforming material between the platens from spreading in
the width direction of the specimen. To ensure that the lateral spread is
negligible, the width of the strip should be 6-10 times the breadth of the
platens, that is, 6 < wfb < 10. The ratio, bft, of the platen breadth to strip
thickness at any instant should be between 2 and 4 and the platens changed
as required during the test to maintain this condition.
If the thickness of the specimen is too great then the indentation pressure
exceeds the yield stress as explained in chapter 7. When the thickness of
the specimen, t, at any instant is infinite then the process becomes an inden-
tation by a single platen for which the indentation pressure is 2.57Y, where
Y is the uniaxial yield stress. If, however, bft is not less than 2 the maximum
error arising from this geometrical constraint is expected to be an overesti-
mate of the uniaxial yield stress not exceeding 2 per cent. This error will be
even less if the ratio, bft, is very large but this implies a large platen breadth, b,
which will tend to increase the frictional resistance. It is necessary to ensure
that the frictional resistance at the specimen-platen interface is minimised
by using a suitable boundary lubricant such as graphite in tallow or molyb-
denum disulphide.
After the specimen has been lubricated and placed accurately between
the lubricated platens in the sub-press, the compressive force is applied
140 ENGINEERING PLASTICITY

by using a suitable testing machine to effect a fractional reduction of about


3 per cent. The compressive axial force is noted and the specimen unloaded.
The strip is then removed from the sub-press and the reduced thickness of
the specimen measured accurately with a micrometer fitted with reduced
anvils which are able to enter the narrow indentation of the specimen. The
specimen is then relubricated, replaced between the platens and the proce-
dure repeated. Incremental loading in the same indentation with relubri-
cation and determination of the compressive strain at each increment is thus
involved.
It is important that, at each increment of loading, the fractional reduction
effected is not less than 3 per cent so that the knee of each stress-strain curve,
corresponding to subsequent yielding, is surmounted as illustrated in
figure 6.10. Otherwise an inaccurate envelope of the stress-strain curves
may be derived or an undue scatter of results obtained.
The ordinate for the incremental test in plane strain compression shown
in figure 6.10 is the mean platen pressure p = F jwb, where F is the axial
compressive force and the abscissa is the compressive natural strain in the
thickness direction ofthe specimen, e1 = ln(t 0 /t).
Assuming that a particular test commences with a prestrain corresponding
to the point A in figure 6.10, then point B is the actual yield point and the
point C is where it first coincides with the envelope of the incremental
tests. The yield stress is defined by extrapolating the envelope to zero strain
corresponding to the point A for the particular test which is then given
by the point D in the diagram. It is therefore necessary that the increment

Envelope of incremental tests

-
.a
it C7'·Yr31/2'2'p P. -----
IL
II

=====1~==--==~::-:.~-----1
p
.,...
Q.
(7'

.,....... 1/?1""..- I . ..
:::J , ............. 1

.,.'8 I DeriVed UniaXIal


.,. ...... T :
.."
Q.
stress-strain curve
c:: ,.-"" I'Knee'of 1
0 lsubsequentl
Ci. :yield curvel
c:: I I
0 I I
~" I
I
:
I
I I

0 A E:t £.. (2/3112JE:t


Compressive natural thickness strain ~=ln(\/tl

Figure 6.10. Typical results from a plane strain compression test


PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 141
of loading be sufficient in each case to ensure that the fractional reduction
achieved is such that the point C is exceeded.
If the specimen material is assumed to be incompressible since the principal
strain increment in the lateral direction, say, dew is zero, the tensile longi-
tudinal principal strain increment, de1 , is equal in magnitude to the com-
pressive thickness strain increment de1 , that is, de1 + dew + de1 = 0 and
de1 = - de1, since dew = 0.
For narrow platens, the principal stress, a 1, in the longitudinal direction
of the strip can be considered zero. The normal platen pressure p = a 3
and the principal stress in the lateral direction a 2. From the Levy-von
Mises stress-strain increment relationship
dew= (2/3)dA[ a 2 - {(a 1 +a 3)/2}] = 0 (6.55)
Therefore a 2 = (a 1 + a 3)/2 = a 3/2
or p=a 3 =2a2 (6.56)
The representative or equivalent stress
jj = (1/2)1/2 {(a1 -a 2f +(a2 -a 3)2 + (a3 -a 1f }1/2
and if the values a 1 = 0, p = a 3 = 2a 2 are substituted then
jj = y = (31/2 /2)p (6.57)
It follows that the mean platen pressure, p, in a frictionless plane strain
compression test is therefore (2/3 1' 2) Y or about 15.5 per cent greater than
it is in the corresponding uniaxial compression test.
The representative or equivalent strain is given by
e= (21/2 /3) {(el- ew)2 + (ew- et)2 + (el- el)2 }1/2
Substituting the values e1 = - e1 and ew = 0 produces
e= (2/3 112)e1 (6.58)
To derive the true compressive stress-compressive natural strain curve
in uniaxial compression from that obtained in plane strain compression it
is necessary to divide the ordinates by 2/3 112 and also to multiply the abscissae
by 2/3 1' 2. Thus a point such asP 1 on the curve for the plane strain com-
pression test becomes the point P 2 on the uniaxial stress-strain curve, as
illustrated in figure 6.10.

6.4 EFFECT OF TEMPERATURE AND STRAIN-RATE


The manner in which the stress-strain relation for a metal is affected by
temperature and strain-rate is, obviously, of importance to engineers in
general. In structural design, plastic deformation is not generally expected
to occur and the range of temperature involved is normally due to climatic
changes. The structural designer is then concerned with how the yield stress
142 ENGINEERING PLASTICITY

of the metal is affected by a relatively small change in temperature and


where the strains are of the order 10- 3 < e < 10- 2 • Nevertheless, to structural
engineers a change in the yield stress ofthe metal by a few per cent, as affected
by temperature and strain-rate, may be of considerable importance.
However, metal forming processes may be performed at high speed and
high temperatures when the flow stress of the workpiece material may be
significantly affected and natural strains of the order of unity are often
encountered. On the other hand, the manufacturing engineer engaged in
metal forming would only become concerned if the effect of strain-rate was
to increase the flow stress of the workpiece material by 50 per cent or more.
Recovery, recrystallisation and grain growth are thermally activated
processes which lead to a reduction in the flow stress, whilst an increasing
rate of plastic deformation of most engineering metals produces an increase
in the flow stress of the metal. It is difficult to generalise on the combined
effects of these two contrary effects of temperature and strain-rate for all
metals because of the metallurgical factors involved, such as phase changes
which can occur especially in highly alloyed metals.
For more than fifty years much effort has been devoted to studying
these phenomena but no clear quantitative relationships have been estab-
lished. Because of the complexity of this field it appears to the author that
no important systematisation will emerge in the near future. However, it is
possible to establish some systematisation of the present knowledge and this
section is presented to provide a background on some important concepts.

6.4.1 Ductile and brittle metals-transition temperature


The behaviour of metals classified as either ductile or brittle has historically
resulted from the performance of the metals according to a quasi-static
uniaxial tensile test. A metal which fractures after very little strain and
reduction in cross-sectional area is classified as brittle and the fracture
exhibited is referred to as a cleavage fracture where the fracture surface
occurs essentially perpendicular to the longitudinal axis of the specimen.
On the other hand, a metal which can be subjected to, say, more than 10 per
cent uniform elongation and where the fracture is typically of a cup and cone
type is classified as ductile and the fracture is referred to as a shear type
fracture.
In these terms, cast iron is regarded as a brittle metal but because of the
free graphite in the crystal structure of cast iron its behaviour can be consider-
ed to be comparable with some non-metallic brittle substances. It should
be appreciated that a so-called ductile metal such as mild steel can exhibit
a brittle fracture when subjected to impact loading. It was shown by Parker,
Davis and Flanigan 17 , in 1946, that by reducing the test temperature suffi-
ciently the shear type fracture normally exhibited by mild steel in uniaxial
tension at ambient temperature can be changed to a cleavage type similar
to the fracture of cast iron at ambient temperature. For some metals including
._.---1---€- 100/min
.----€= 19000/min

Figure 6.11. The effect of temperature and strain-rate on the ductility of body-
centred tetragonal tin in uniaxial tension

mild steel, molybdenum, chromium and tin there is a transition temperature


above which the metal is ductile and below which it is brittle. A compre-
hensive review of the low temperature brittleness of metals is given by
Tipper 18 • Metals which can be classified as ductile but do not exhibit a
transition temperature and can therefore be plastically deformed at all
temperatures include aluminium, copper, nickel, gold, silver, platinum
and most of their alloys. Magnesium and its alloys, tungsten and molyb-
denum, are insufficiently ductile to be formed at ambient temperature.
Those metals which have a marked transition temperature are more generally
those having a body-centred cubic structure whilst those which do not
exhibit a transition temperature are generally those having a face-centred
cubic structure.
The transition temperatures for body-centred tetragonal tin in uniaxial
tension corresponding to different strain-rates were originally determined
by Magnusson and Baldwin 19 and are illustrated in figure 6.11. This figure
shows, for a given strain-rate, that there is a small range of temperature
below which the percentage reduction in area is small and approximately
constant when the metal is brittle and above which it is high and nearly
constant when the metal is ductile.

6.4.2 The significance of recrystallisation and homologous temperatures


It has been customary to refer to cold forming when metal is processed
at room temperature·and the metal is permanently strain-hardened during
the forming process. Although most engineering metals strain-harden at
room temperature, pure lead, tin and cadmium only permanently strain-
harden at temperatures below room temperature and, after straining, these
metals will ultimately recrystallise at about room temperature if given a
sufficient period of time.
Above the appropriate recrystallisation temperature say, TR, which is
144 ENGINEERING PLASTICITY

known to depend on the strain and strain-rate, a metal can be subjected


to plastic deformation when the flow stress remains sensibly constant due
to the rate of strain-hardening being nullified by the rate of thermal recovery.
The flow stress, however, may be highly strain-rate sensitive. The effect of
temperature during the quasi-static compression (6 < 10- 3 /s) of commer-
cially pure aluminium (B. S. 1470 SIC), copper (B.S. 1432) and En 2 steel 20
is illustrated by the true compressive stress-natural compressive strain
(u,e) curves for 0 < e < 0.8 given in figures 6.12, 6.13 and 6.14, respectively.
In each case, it can be seen that, at room temperature, the metal exhibits a
strain-hardening characteristic and as the temperature increases the rate
of strain-hardening decreases. At some temperature, the curves for each
metal first become flat, that is, the rate of strain-hardening is nullified
by the rate of thermal recovery. This corresponds approximately to the
recrystallisation temperature of the metal. The true compressive stress-
natural compressive strain curves shown in figure 6.15 are for En 52 steel
at a constant temperature of 1000°C for various natural strain-rates (1.5/s <
6 < 100/s) which were obtained by Cook 21 from axisymmetric compression
tests employing a cam plastometer. The cam plastometer is a straining
machine in which a cam having a J.ogarithmic profile actuates a moving
lower platen to compress a specimen against a stationary upper platen at a
constant natural strain-rate. The cam plastometer was originally developed

140.------.------..------....-------,9

200°C

250°C

300°C
"
""
L.
1-
500°C

0•8

Figure 6.12. Quasi-static, true compressive stress-natural compressive strain


characteristic curves for aluminium (B. S. 1470 SIC) at elevated temperatures (after
Slater and Johnson 20 )
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 145

~400 N
25 -~
~ ~
b ...c
0

....... ....
:1300

..
...
!;
>

...
·:200
a. 400°C 10
.?:
S
E ~~~------------~~~1 ~

...
0
u E
0
----------~==.,r-~5 ~
.
:::J
1- :::J
1-

0 0·8

Figure 6.13. Quasi-static, true compressive stress-natural compressive strain


characteristic curves for copper (B.S. 1432) at elevated temperatures (after Slater and
Johnson 20 )

.
.
>
IS·=
a.
E
0
u

0·2 0•4 0•6 0·8


Natural compressive strain E:
Figure 6.14. Quasi-static, true compressive stress-natural compressive strain
characteristic curves for En 2 steel at elevated temperatures (after Slater and Johnson 20 )

by Orowan 22 to obtain stress-strain data at constant natural strain-rates


for purposes of rolling mill force and torque estimation. For further discus-
sion of the subject generally the textbook by Bell 23 should be consulted
and an account ofthe behaviour of high strength metals at cryogenic tempera-
tures is given by Coffin and Conrad 24 •
146 ENGINEERING PLASTICITY

E:•IOO/s

.. E:= 40/s

..."...
..
~200
"iii
...
"
0..
E
0100
u
5 ~
"...
:::J ...
1-
1-

0 C>-2 0·4 0·6 0·8


Natural compressive strain E:

Fig11re 6.15. True compressive stress-natural compressive strain characteristic


curves for En 52 steel at a temperature of l000°C and various natural strain-rates
(after Cook 21 )

The interpretation of stress-strain-temperature-strain-rate data for


metals and the comparison of experimental results obtained by different
investigators is best facilitated on a dimensionless basis. It is preferable
to reduce temperatures to the dimensionless homologous temperature, T",
which is defined as the absolute test temperature, T, expressed as a fraction
of the absolute melting point temperature, TM, of the test material where
the absolute temperatures are determined, say, on the Kelvin scale, that is
(6.59)
Comparisons of data for metals having different melting point tempera-
tures are then likely to be more meaningful. It is found that the absolute
recrystallisation temperature for metals, TR, expressed in terms of the
homologous temperature is given by
0.4 < TR < 0.55 (6.60)
but is dependent on the strain and strain-rate. However, the statement
(6.60) is a useful guide.
The significance of the recrystallisation temperature in the interpretation
of data is well illustrated by figures 6.16(a) and (b). These figures show
some of the results obtained by Mahtab, Johnson and Slater25 when using
a lubricated, hardened tool steel, cylindro-conical indenter of 90° conical
angle in a quasi-static test to indent annealed copper (B. S. 1433) and an
annealed aluminium alloy (B. S. 1476 HE 10). It should be noted that an
indentation test is indicative of the stress-strain relation for a material
because the mean indentation pressure corresponds to the mean yield stress
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 147

8 Hiatus

\,j
0
a.
::?! 6
400
O."' 200
I'
"'c
...
" '' "'
I
c
5300 I
'' "-
.... "....
.
I
4 lr ' \ \I
t\ .a 15
100

ro
lu
' ' ..,o
.a
I \ \I
I \
",, "'0
'' '
t 10
~200 "I"'
ol<'> 3
I
\
a
..
·;;I 8 )( I \ )(
0 50 I
;; "I ..
~~~ a."' '' a."' 6
0
01 ...
c ~~~
"
'0
.!:
>-I ..
bl~ 1 1 4
a::,-
.. I ..
20

'
c Annealed
~ 100 aluminium alloy 2
::?! Annealed copper
(6.5.1433) 10 (6.5.1476 HE 10)
I
0 800 0

Figure 6.16. The significance of the recrystallisation temperature in relation to


change of mechanical properties: (a) annealed copper (B.S. 1433); (b) annealed
aluminium alloy (B.S. 1476 HE 10) (after Mahtab, Johnson and Slater 25 )

for a given mean strain. Figure 6.16 shows the relation between the logarithm
of the mean quasi-static indentation pressure and the temperature for the
two metals. Referring to figure 6.16(a) for the annealed copper it will be
observed that there is a discontinuity of slope and for the annealed aluminium
alloy there is a hiatus in figure 6.16(b) which occur in each case at approxima-
tely the corresponding recrystallisation temperature.

6.4.3 The ratio of dynamic to quasi-static flow stress


When presenting data on a dimensionless basis, it is also useful to determine
how the ratio of the dynamic stress, aD, to the quasi-static stress, as, of a
metal for a particular strain varies with the homologous temperature, Tu.
For example, during a slow speed axisymmetric compression test, the
quasi-static flow stress, as, is determined at a natural strain-rate e < 10- 3 js
and the dynamic flow stress, aD, would be determined from the results of
some form of impact test where natural strain-rates of the order 102 /s may
be encountered. In this case, the range of strain-rate ratio involved is then
10 5 . Figure 6.17 shows the relation between the stress ratio, aD/as, and
the homologous temperature, TH, for 0.55 per cent plain carbon steel 26
for a constant compressive engineering strain e = 0.15 and for various
compressive engineering strain-rates.
This kind of relationship is typical of the behaviour of a binary alloy.
148 ENGINEERING PLASTICITY

Approximate recrystallisation
temperature of a iron
(ferrite I

I
Approximate
I
1 A 3 phase
:---- transformation
: la--y)
1 temperature
I
I

1·0
Homologous temperature TtrT/TM
Figure 6.17. Relation between the stress ratio (u0 /u 5) and homologous tempera-
ture (TH) for 0.55 per cent plain carbon steel at a constant compressive engineering
strain of 0.15 for various compressive engineering strain-rates (after Slater, Ak:u and
Johnson 26 )

It will be observed that inflexions occur in the curves corresponding to


(a) the approximate recrystallisation temperature of the primary constituent
of the alloy which, for the steel considered, is tx iron (ferrite), and (b) the A3
phase transformation temperature when the body-centred cubic structure
of the tx iron (ferrite) changes to the face-centred cubic structure of y iron
(austenite). The relationship for pure metals with no allotropic changes
will only exhibit one inflexion corresponding to the recrystallisation tempera-
ture. It will also be observed from figure 6.17 that when Ts. ~ 0.8 and e =
300/s, u0 /u8 ~ 8.
It can thus be concluded, when the strains are large, say between 0.1 and
1, that:

(1) The effect of strain-rate is not significant at homologous temperatures,


Tn, less than about 0.5, that is, less than the recrystallisation temperature,
~.and 1 < u0 /us < 2 approximately.
(2) The ratio, u0 /u8 , is very sensitive to strain-rate at temperatures above
the recrystallisation temperature, TR, say, TR > 0.6 and for some metals this
ratio is known to be greater than 10.

These facts do not appear to be generally recognised at present. The latter


point is very important because when a metal is deformed at temperatures
exceeding the recrystallisation temperature and at high speed its resistance
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 149
to deformation or stiffness is greatly increased compared with the quasi-
static situation. As figure 6.17 shows, the high speed deformation of 0.55
per cent plain carbon steel at temperatures in excess of T" = 0.4 is more
difficult than the slow speed deformation at room temperature.
There are certain economic advantages to be gained in processing some
metals at temperatures which are intermediate between room temperature
and the recrystallisation temperature, for example, the extrusion of steel
bar. Metal forming in this range of temperature has become known as
warm forming. It is now possible to suggest that the distinction between
cold forming, warm forming and hot forming can be made with reference to
the recrystallisation temperature, TR, and the homologous temperature, 4I,
as follows : cold forming, T" < TR ~ 0.4; warm forming, 0.4 < TH < 0.6;
hot forming, T" > 0.6.
Numerous investigations have been performed to determine stress-
strain-temperature-strain-rate data for various metals employing a variety
of experimental techniques. These different test techniques are briefly
reviewed in the textbook by Johnson and Mellor 2 7 • Representative of these
investigations are those by Alder and Phillips 28 , Manjoine and NadaF 9 ,
Manjoine 30 , Samanta 31 • 32 , Malvern 33 and, more recently, by Dean and
Sturgess 34 . However, the results of these studies have not been presented in
the dimensionless form previously discussed.

6.4.4 Strain-softening
During plastic deformation of a metal, external work is done and this
dissipation of energy is ultimately transformed into thermal energy so that
an increase in the temperature of the metal may occur. At temperatures
below the recrystallisation temperature and during slow speed forming the
effects of this increase in temperature do not appear to be significant. How-
ever, it is logical to assume that, for particular metals, at some condition of
temperature and strain-rate, with this contribution to the thermal energy
possessed by the metal, the rate of thermal recovery will exceed the rate
of strain-hardening. In this event, the flow stress of the metal would be
expected to decrease with increasing strain, that is, strain-softening would
occur. Indeed, strain-softening is exhibited in the characteristic curves of
figure 6.15 for strains, B > 0.3.

6.4.5 Yield stress of steel at about room temperature and yield delay
From quasi-static and impact tension and compression tests in the range
of strain-rate, 10- 3 /s < 8 < 103 /s, as performed by Hopkinson 35 , Campbell
and Duby 36 •37 and many others it may be concluded that the ratio of the
dynamic stress to the quasi-static initial yield stress for mild steel is usually
about 2 and sometimes as high as 3. However, if instead of initial yield
stress, the flow stress is considered for the comparison then this ratio is
150 ENGINEERING PLASTICITY

generally less and decreases with increasing strain. As stated previously,


it is also found that the flow stress for a given strain generally increases
with strain-rate. The stress ratio relating the initial yield stress for mild steel
decreases if the time available for loading increases, for example, the ratio
decreases from about 2.5 to 1.3 when the time increases from 10- 5 to 10- 2 s.
Furthermore, it appears that the ratio is smaller the higher the actual value
of the quasi-static initial yield stress.
In the forming of annealed low carbon steels there is a perceptible delay
time from the instant of loading to the onset of plastic deformation. This
phenomenon is associated with the existence of a distinct yield point and
yield mechanism which was discussed in sections 6.2.1 and 6.2.4.
It appears that the time required to yield decreases as the loading increases.
The maximum delay time is about 1 sat a temperature of 25oC but if the delay
time is reduced to, say, 10- 6 s the ratio of the dynamic to quasi-static initial
yield stress is about 2.5. Assuming this ratio to remain unaltered then, at a
temperature of - 60°C, the delay time may increase to over 10 s. At a tempe-
rature of 121 oc the delay time is reduced to between 10- 2 and 10- 3 s. At
higher temperatures, the delay time does not decrease rapidly. The textbook
by Goldsmith 38 presents an extensive summary of the work concerned with
these topics.

6.4.6 Correlation between stress, strain-rate and temperature


In 1909, Ludwik 9 proposed a semi-logarithmic relation between tensile
stress and strain-rate. The equation has the form
(6.61)
where a 1 , a 0 and £0 are constants. Constant strain-rate tests were performed
to obtain (a,a) curves for a given strain-rate and temperature. It appears
that equation (6.61) is valid for a number of metals for a certain range of
strain and provided that the tensile stress, a, does not increase rapidly with
strain-rate t.
Alder and Phillips 28 conducted experiments on the dynamic compression
of aluminium at temperatures between 20°C and 550°C, copper up to
temperatures of 600°C and on steel up to l200°C. The range of natural
strain-rate was 1 to 40/s. These authors summarised their results in terms
of a power law given by
(6.62)
where a 0 and n are constants. It was concluded that equation (6.62) gave
a better interpretation of their results than did the Ludwik-type equation
(6.61). However, for equation (6.62) to be dimensionally valid, it is preferable
to replace it by
(6.63)
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 151
where u 0 is the flow stress corresponding to a strain-rate, 80 , at a constant
strain.
If the data obtained by Alder and Phillips are analysed for compressive
strains up to 0.5, the relation between the index n and the homologous
temperature can be determined. These results show that the mechanical
properties of the metals change at about the recrystallisation temperature
for each material so that when
TR < 0.55, n l't; 0.055
and when TR > 0.55, n l't; 0.43
The concept of a velocity modified temperature, Tmod• was introduced by
Macgregor and Fisher 39 in 1946. An approximate relationship was proposed
which is given by
(6.64)
and the velocity modified temperature depends on the absolute temperature,
efe0 , where eis the actual equivalent
T, a constant k and the strain-rate ratio
strain-rate and 80 is that of some standard state. Substitution of 0( = k/
(1 + k lne0 ) into equation (6.64) produces a form of equation in which the
standard state is incorporated in the constant 0( so that
ii = f(Tmod) = f { T(1 - 0( In~)} (6.65)
It should be noted that equation (6.65) demonstrates the expected qualita-
tive result that an increase in strain-rate is equivalent to a decrease in tem-
perature. However, at temperatures less than the recrystallisation tempera-
ture, equation (6.65) can only be regarded as approximate.
The following equation involving strain, strain-rate and temperature
was introduced by Inouye 40 in 1955
(6.66)
where u 0 , n, m, A and k are constants.
From his experimental studies concerned with annealed aluminium,
Malvern 33 introduced the equation having the form
e= (u/E) + F {u- Uo(e)} (6.67)
where u > u0 (e), u 0 (e) is the quasi-static stress at a strain, e, and E is Young's
modulus of elasticity.
Other forms of Malvern's equation are
"e=(u/E)+D{(ufu0 )-1}P (6.68)
and e= (u/E) + A[exp{(u/u0 ) - 1}q- 1] (6.69)
where A, D, p and q are empirical constants and u 0 is the quasi-static yield
stress.
Equation (6.67) predicts an increase in the initial yield stress at high
152 ENGINEERING PLASTICITY

strain-rates and also permits plastic strain increments to be propagated


at elastic wave speeds along bars which are initially stressed into the plastic
range 41 .
Malvern's equation was adapted by Ripperberger42 in 1965 to summarise
his experimental data obtained for annealed alummium. This equation has
the form
(6.70)
where r is the relaxation time of the material, u 0 (e) is the quasi-static stress
at a strain e and, therefore, u- u 0 (e) is the excess stress to which the material
is subjected at a given strain-rate.
As a result of his dynamic wave propagation studies, Bell suggested that
the following stress-strain function has wide and remarkable generality
(6.71)
where Jl{O) is the zero point isotropic linear elastic shear modulus, B0 = 0.0280
is a dimensionless universal constant and r is an integer. Bell claims that
equation (6.71) is valid for the entire temperature range from 4 oK to within
20° of the melting point of crystalline solids and for strain-rates 10- 9 /s <
e < 104 /s.
Records of the nonlinear wave front generated during uniaxial impact
experiments, after the front had travelled one specimen diameter from the
impact face, were used to determine the stress-strain function

u =I: p 0 c;(e)de (6.72)

where p 0 is the density of the metal, and cP(e) is the plastic wave speed. A
diffraction grating technique was used by Bell to determine large strain
in microsecond time.
Further discussion of this topic and some other equations correlating
stress, strain-rate and temperature are given in the textbook by Thomsen,
Yang and Kobayashi 43 .

6.5 ANISOTROPY AND AELOTROPY


An engineering metal is a polycrystalline aggregate with crystal grains of
various shapes and orientations. If the orientations are randomly distributed,
and the average dimensions of the individual crystals are very small compared
with the dimensions of the specimen, then the metal is considered to be
macroscopically isotropic. However, the properties of an aggregate are
not always given as the statistical averages of the properties of a single
crystal for the different orientations. Although this may be approximately
true for those properties which depend mainly on the bulk structure such
as the elastic moduli it is not necessarily true of plastic deformation pheno-
mena.
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 153
When a metal is stressed such that plastic deformation occurs producing
a large distortion of one kind as, for instance, during extrusion, drawing
or rolling, the crystallographic directions in each grain are gradually rotated
towards a common axis. A preferred orientation is thus created so that in
the case of a metal in which the grains were initially orientated at random,
that is, an isotropic material, it becomes increasingly anisotropic when
plastically deformed. This implies that the yield stress and ductility of the
metal are dependent on the direction in which they are measured.
A special case of anisotropy in which the properties are equal in two
orthogonal directions is known as aelotropy. Probably, the most familiar
material which is aelotropic is wood. Due to its grained structure produced
by the growth process of the original tree it is stronger when stressed along
its grain than when stressed across the grain.

6.5.1 Planar and normal anisotropy in sheet metal


In sheet metal products, anisotropy is often considered only in a two-
dimensional sense. A difference in properties measured in various directions
within the plane of the sheet is referred to as planar anisotropy. However,
differences occur between properties measured normal and parallel to
the plane of the sheet which are termed normal anisotropy. Absence of
anisotropy in the plane of the sheet does not assure that the flow strength
perpendicular to that plane is the same. Normal anisotropy of, for example,
low plain carbon steel sheet is sometimes more pronounced than its planar
anisotropy and is often desirable for good drawability, whereas, planar
anisotropy is usually undesirable.

6.5.2 Strain ratio or 'R' value


Although it is difficult to measure the flow strength of sheet metal in its
thickness direction directly, the degree of planar and normal anisotropy
can be assessed from the results of uniaxial tensile tests performed on
specimens cut from the sheet in different directions, say, relative to the
direction of rolling of the sheet. It is then found that for a given elongation,
different strains are exhibited in the width and thickness directions of the
specimens. The ratio of these strains, however, is sensibly constant and is
referred to as the strain ratio or 'R' value which is defined as
(6.73)
where ew is the natural width strain and e1 is the natural thickness strain.
The R values of sheet metals are usually assessed for longitudinal strains,
e1, greater than about 0.05. It can then be assumed that the elastic strains
are negligible and the metal can then be considered as incompressible so
that the algebraic sum of the principal natural strains is zero, that is,
(6.74)
154 ENGINEERING PLASTICITY

Hence e1 = - (e 1 + ew) (6.75)


It should be noted that e1 = ln{lj/ 0 ) is tensile and positive whilst ew =In
(w/w 0 ) is compressive and hence negative, where /0 is the initial gauge length;
l the current gauge length after a specified elongation of the specimen;
w0 the initial mean width of the specimen; and w the current mean width
after a specified elongation of the specimen.
The measurement of 'R' value has been considered in some detail by
Atkinson 44 .
Normal and planar anisotropy can be assessed by means of two para-
meters which assist in defining the deep-drawing behaviour of sheet
metals 45 •46 • Normal anisotropy is assessed by the average strain ratio, R, and
the degree of planar anisotropy is reflected by the variation in strain ratio,
AR, as illustrated in figure 6.18. An average strain ratio of unity indicates
equal flow strengths in the width and thickness directions of the sheet and
such a metal is regarded as possessing only fair drawability. The ideal
deep-drawing material for the manufacture of symmetrical products is
one which is not only ductile but has both a high average strain ratio,
R > 1, and a AR value of zero.
If the strain ratio in the longitudinal and transverse directions is greater
than in the 45° direction then AR is considered positive as shown in figure
6.18(a). A negative value of AR, as shown in figure 6.18(c), indicates that
the strain ratio in the 45° direction relative to the direction of rolling is
greater than in the longitudinal and transverse directions.
Assume that a strip of sheet metal is contained in the (x, y) plane and is
cut from the sheet with its longitudinal axis parallel to the x axis of anisotropy.
From equations (5.80), the incremental plastic strain ratios are then given by
de~ :de~ :de~=G+H: -H: -G (6.76)

(a) (b) (c)

.
2 2r- 2

-"'
Ill

II
a:
ol II-
..
·.;::;
c

....
c:
c
II)

0 45
Orientation to the direction of rolling
Figure 6.18. Variation of strain ratio with orientation
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 155
The ratio of the width strain to the thickness strain or Rx value is
Rx = de~/de~ = H/G (6.77)
where the subscript x denotes that the specimen is orientated along the
x direction. For a strip cut from the sheet in they direction
de~ :de~ :de~= -H :F+H: -F (6.78)
and R y = dePfdeP
X Z
= H/F (6.79)
To derive the anisotropic parameters for plane stress deformation of
a sheet metal it is necessary to perform a uniaxial tensile test in at least
one other direction in the plane of the sheet. It may be assumed that if the
specimen is subjected to stress in the plane of the sheet, that is, the (x, y)
plane then the shear stresses 'tyz and 'tzy are zero. If a specimen is cut with
its longitudinal axis at an angle a to the x direction, then
ax= a cos 2 (X
aY = a sin2 a (6.80)
and 'txy = a sin (X cos (X
where a is the applied tensile yield stress and, if these values are substituted
into equations (5.80), the following equations are obtained
de~= {(G +H) cos 2 a- H sin 2 a}adA.
de~= {(F +H) sin 2 a- H cos 2 a}adA. (6.81)
dt~~ = - {F sin 2 a + G cos 2 a} adA.

and rly~Y = (N sin a cos a)adA.


The width strain increment de~+(n/ 2 > is given approximately as
d e,.+(n/2)
p
= dP"2
ex sm (X+ dP
By cos 2 (X- 2dP.
Yxy sm (X cos (X (6.82)
Therefore R,. = de~+(n/ 2 /de~
= (de~ sin 2 a + de~ cos 2 a - 2dy~Y sin a cos a)/de~
= {H + (2N- F- G- 4H) sin 2 a cos 2 a}/(F sin 2 a+ Gcos 2a)
(6.83)
The direction of rolling of a sheet metal is usually an axis of anisotropy
in which case the x direction can be assumed to coincide with the rolling
direction.
Equation (6.83) then reduces to

}
Rx=R 0 =H/G
Ry=R 90 =H/F (6.84)
and R 45 = {N - !(F + G)} /(F + G)
Therefore N /G = {R45 + (1/2)} {1 + (R 0 /R 90 )} (6.85)
156 ENGINEERING PLASTICITY

where R 0 , R 45 and R 90 represent the R values in the direction of rolling,


at 45o and 90° or transverse to the direction of rolling respectively. The
derivation of the strain ratios given by equations (6.84) assumes that the
anisotropic parameters remain constant. This assumption should be verified
experimentally.

6.5.3 Effects of anisotropy during sheet metal forming


Normal anisotropy present in sheet metals often imparts good drawability
whilst planar anisotropy usually produces undesirable features. For example,
during the deep drawing of a symmetrical component such as a cylindrical
cup, if the strain ratios in the longitudinal and transverse directions of the
sheet metal are greater than in the 45° direction as is shown in figure 6.18(a)
then earing will occur in the 0 and 90° directions. If the sheet exhibits a
negative !'lR value, then ears are formed in the direction which is at 45°
to the direction of rolling. The variation of the strain ratio, !'lR, thus provides
a good indication of the earing behaviour of the material because ears
are usually formed in the directions in which the strain ratio is a maximum.
It follows that the greater the value of !'lR, the more pronounced is the
earing. Absence of earing occurs when !'lR is zero. This earing feature
experienced during the deep drawing of a cylindrical cup is undesirable
when a uniform rim is required. A secondary machining operation is necessi-
tated thus increasing manufacturing costs.

6.5.4 Control of anisotropy in low plain carbon steel sheet


Sheets of low plain carbon steel exhibit various degrees of anisotropy
depending upon the manufacturing history. The extent to which the steel
is cold-rolled prior to annealing, the hot-mill finishing and the final coiling
temperatures are known to influence the final anisotropy of the sheet steel.
A low plain carbon steel has a body-centred cubic structure as shown in
figure 6.19(a) and has greatest resistance to deformation in the direction
of the cube diagonal (111 ], less resistance along the face diagonal [110]
and least along the cube edge [1 00]. If the crystals have a preferred orientation
due to plastic deformation during a sheet metal forming process, then the
bulk material will tend to exhibit the anisotropic characteristics of a single
crystal. It was shown by Burns and Heyer 4 7 that a sheet with a 'cube-on-
corner' texture, in which the cube diagonal is orientated perpendicular
to the plane of the sheet, should develop high strain ratios. Figures 6.19(b)
and (c) are for aluminium-killed steel sheet, cold-rolled to various fractional
reductions. It will be noted that the strain ratio increases with intensities
of 'cube-on-corner' [111] texture. However, the average strain ratio, R,
decreases for fractional reductions greater than about 0.75 whilst the inten-
sity of the 'cube-on-corner' texture continues to increase until the fractional
reduction reaches 0.9. Changes in the average strain ratio are more closely
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 157
z

.
II%
.2 2·0r---.----.------.----.-----,
!:!
-6. 1·5
...
"'~1·0
!:!
"
~0·5'-----'------...l.-----'--------L----'

lbl
4 Intensity of [Ill] peak
• Reciprocal intensity
of [I 00] peak
Absence of

:_..Aio---- cube-on-corner
[Ill] texture

0 0•4 0·6 0•8


fractional reduction
(c)

Figure 6.19. (a) Body-centred cubic structure of a low plain carbon steel;
(b) variation of the average strain ratio, R, with cold fractional reduction; (c) variation
of final annealed texture with cold fractional reduction

reflected by c}\anges in the 'cube-on-face' [100] texture but in an inverse


manner. Since the presence of grains with a 'cube-on-face' texture is expected
to lower the average strain ratio, the absence of such grains is expected
to raise the strain ratio.

6.6 BAUSCHINGER EFFECT


In general, the effect of plastically deforming a metal at room temperature
is to increase its resistance to further deformation by virtue of strain-harden-
ing. This provides a means of improving the strength and hardness of a
component and there are many applications where this property is utilised
to advantage. These include shot peening, cold rolling and also the auto-
frettaging of a thick-walled cylinder to increase its elastic limit for reloading.
However, such applications assume that only a monotonic stress is involved.
158 ENGINEERING PLASTICITY

+U

-u
Figure 6.20. The Bauschinger effect

When a metal is stressed to produce plastic deformation and then un-


loaded, residual stresses on a microscopic scale remain due mainly to the
different states of stress existing in the differently orientated crystals before
unloading occurs. If a reversal of stress now occurs then such residual
stresses could be expected to have some influence on plastic yielding. Suppose
that a specimen is subjected to a uniaxial tensile stress which exceeds the
initial tensile yield stress, + Y, so as to produce plastic deformation, corres-
ponding to point A in figure 6.20 and then unloaded to point B. Neglecting
hysteresis, the unloading will occur elastically and an irrecoverable plastic
deformation results. On reloading in tension, the reloading path follows
the elastic line BA and the subsequent tensile yield stress, + Y1 , is greater
than the initial yield stress + Y. If, however, the specimen after unloading
to the point B is now subjected to uniaxial compression it is observed that,
because of residual stresses present on unloading, yielding of the specimen
as a whole occurs at a reduced magnitude of stress, - Y2 , and it is possible
that this may even be lower than the initial tensile yield stress, + Y.
This phenomenon is known as the Bauschinger effect 48 and is present
whenever a reversal of stress occurs. Since the effect is known to be absent
from single crystals of pure metals it is believed to be attributable to a
particular kind of residual stress influenced by the grain boundaries. It
would therefore appear that components which are to be, say, subject to
tension in service should not be strain-hardened by compressive loading.
The explanation given here for the existence of the so-called Bauschinger
effect is that which is at present generally accepted, but another explanation
PHENOMENOWGICAL NATURE OF ENGINEERING METALS 159
based on the anisotropy of the dislocation field produced by loading is
given in the textbook by McLean 49 • Several simplified models have also been
proposed to describe the Bauschinger effect as, for example, given by Goodier
and Hodge 50 •
The residual stresses and consequently the Bauschinger effect can be
removed by a low temperature heat treatment. In contrast, to change the
preferred orientation of crystals responsible for the anisotropy of a metal
it is necessary for the heat treatment to be carried out above the recrystallisa-
tion temperature.

6. 7 PLASTIC INSTABILITY

6. 7.1 Effect of elastic deformation on instability in uniaxial tension


Plastic instability in uniaxial tension was considered in section 6.2.6 and
if the stress-strain relationship for a strain-hardening material is known
it was shown that the instability strain can be estimated. The solution is
now extended to take into account the effect of elastic deformation.
It may be assumed that the longitudinal direction is the first principal
strain direction and the corresponding total strain increment is de 1 • The
plastic strain increment for this direction is
de~ =de 1 -(datfE) (6.86)
and for the transverse directions
de~= de 2 + v(datfE) (6.87)
and de~= de 3 + v(da 1 /E) {6.88)
where E is Young's modulus of elasticity and v is Poisson's ratio.
For an isotropic material de~ = de~ and if the material is incompressible
de~+de~+de~=O

Therefore
de~= de~=- deV2 =- {de 1 - (datfE)}/2
or de 2 + v(datfE) = de 3 + v(datfE) =- {de 1 - (datfE)}/2
and de 2 = de 3 =- {de 1 - (1- 2v)(da 1 /E)}/2 (6.89)
If there is no rotation of the principal axes it follows that
e2 =e 3 = -{e 1 -(l-2v)(atfE)}/2 (6.90)
The tensile axial force F, at any instant before the condition of instability
is reached is given by
F = a 1 A = a 1 A 0 /(1 + e 1 )
where A 0 and A are the initial and current cross-sectional areas respectively
160 ENGINEERING PLASTICITY

and e 1 is the principal engineering strain in the longitudinal direction.


Therefore
F =a 1 A 0 exp( - 81 ) =a 1 A 0 exp(28 2 ) (6.91)
smce 81 = ln(1 + e1 ) = -28 2
Combining equations (6.90) and (6.91) produces
F=a 1 A 0 exp[- {~: 1 -(1-2v)(a 1 /E)}] (6.92)
If there is no elastic volume change, that is, v = 1/2, the condition for instabi-
lity in uniaxial tension is da dd8 1 =a 1 which agrees with equation (6.43).
However, when the elastic volume change is not neglected, the condition
for instability in uniaxial tension is then given by
dF/d8 1 = 0 =- a 1 + (dadd~: 1 ) + {a 1 (1- 2v)/E}(dadd8 1 )
or da dd8 1 =a 1 /[1 +{a 1 (1 - 2v)j E}] =a 1 / {1 +(a 1 /3K)} (6.93)
where a 1 is the true stress at instability. It should be noted that the term
a d3K is likely to be very small compared with unity. The phenomenon
of instability which occurs in uniaxial tension can also occur when the
material is subjected to complex tensile stress states, and is of importance
when sheet metal is stretch formed or strained in biaxial tension where
the stress in the thickness direction, that is, normal to the plane of the sheet,
is either zero or may be considered negligible compared with the stresses
in the plane of the sheet.

6.7.2 Instability in biaxial tension


For biaxial tension, the principal stress state is a 1 , a 2 , a 3 = 0. Let a 2 = xa 1
where x is the stress ratio and is a proper fraction. The Levy-von Mises
stress-strain increment relations (5.34) referred to the principal directions
are

Therefore
d8 1 /(a 1 - am)= d8 2 /(a 2 - am)= d8 3 /(a 3 - am)
where the hydrostatic stress, am,=(a 1 +a 2 +a 3 )/3=a 1 (1+x)/3 since
a 3 = 0. Hence
de 1/[a 1 - {a 1 (1 +x)/3}] =d8 2 /[xa 1 - {a 1 (1 +x)/3}] = d8 3 /{ -a 1 (1 +x)/3}
which reduces to
d8d(2- x) = d8 2 /(2x - 1) = - d8 3 (1 + x) (6.94)
Similarly, equation (2.46) for the equivalent stress, a, can be rewritten as
a= a 1 (1 - x + x 2 ) 112 (6.95)
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 161

Figure 6.21. Generalised instability strain

The equivalent strain increment, df., referred to the principal directions is


de= [2{ (de 1 - de 2 ) 2 + (de 2 - de 3 ) 2 + (de 3 - de 1 ) 2 } ] 112 /3 (6.96)
and can be rewritten as
(6.97)
Combining equations (6.94), (6.95) and (6.97) produces in the integrated form
e/2(1 - x + x 2 ) 112 = etf(2- x) = e2 j(2x- 1) = - e3 /(1 + x) (6.98)
Assume the sheet metal to be strain-hardening and the equivalent stress-
equivalent strain characteristic curve is approximated by the Swift-type
equation (6.32), that is,
(6.99)
where 0:::::; q :::::; 1 and C, D and q are ccnstants for a given material. The
constant C is some measure of the yield stress of the metal independent
of its initial state; D is an indication of the degree of prestrain; and q is a
measure of the rate of strain-hardening and for engineering metals usually
has a value between 0.2 and 0.5. Differentiating equation (6.99)
dO: fde = qu/(D +e)= o= jz (6.100)
where z = (D + B)jq is the sub-tangent shown in figure 6.21. It was previously
shown in section 6.2.6 that for uniaxial tension z = 1, as illustrated in figure
6.5(a). The magnitude of the constant C does not affect the strain to which
the metal can be subjected. However, the true stress required to attain any
particular value of strain is affected.
162 ENGINEERING PLASTICITY

6.7.3 Hydrostatic bulge test


The hydrostatic bulge test is of importance in determining the stress-strain
characteristic curve for a sheet metal in approximately balanced biaxial
tension implying that the principal stresses are a 1 = a 2 which are both
tensile and a 3 = 0. The maximum strain that can be achieved before the
condition of instability in tension occurs is found to be greater than that
for uniaxial tension.
Consider a thin circular diaphragm or blank of sheet metal of initial
thickness, t 0 , to be clamped at its periphery on to a circular orifice-type
die of throat diameter, 2a, and subjected to a hydrostatic pressure, p, on
one side causing the blank to be deformed into a domed shape of polar
height, h, as shown in figure 6.22. The thickness to diameter ratio of the
blank is assumed to be such that bending and shearing stresses are insignifi-
cant. The shape of the deformed blank is approximately spherical except
in the region where the blank is clamped on to the die. Divergence from
a true spherical shape is reflected in a changing stress ratio except near the
pole where x = 1 throughout for an isotropic material.
The deformed blank can be likened to a thin spherical shell subjected
to internal pressure p. The membran~ stresses are circumferential stresses
which are tensile and the stress state at the pole is a 1 = a 2 = a and a 3 = 0.
The circumferential stresses at the pole are then given by

(6.101)
where p is the fluid pressure, p the radius of curvature of the deformed blank
and t is the current thickness of the blank. Equation (6.101) implies that
rotational symmetry exists about the pole and hence it is assumed that the
deformation is isotropic.
Deformed circular blank Pole Clamp ring

Circu Ior orifice


type die

Figure 6.22. Geometry of the deformed blank during a hydrostatic bulge test
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 163
The equivalent stress, ii, in terms of the principal stresses is
ii = [(1/2){ (a 1 -a 2) 2 +(a 2 -a 3 ) 2 +(a3 -a 1f} ]112
Substituting the values a 1 = a 2 = a and a 3 = 0 yields
a=a (6.102)
thus indicating that the equivalent stress is equal in magnitude to one of
the principal circumferential stresses and, of course, is tensile.
It is assumed that the membrane strains 8 1 = 8 2 = 8 in all directions and
are tensile. For the condition of incompressibility
81 + 82 + 83 = 0
Therefore (6.103)
where 8 3 is the thickness strain given by
83 = ln(t/t0 ) = -ln(t0 /t) (6.104)
The equivalent strain, e, referred to the principal directions is
e= [2{ (81 - 82) 2 + (82 - 83) 2 + (83 - 81) 2 } ]1 12 /3
Substituting the values 81 = 8 2 = 8 and 83 =- 28 results in
e= 28 = - 83 = ln(to/t) (6.105)
showing that the equivalent strain is equal in magnitude to the thickness
strain but is tensile, as would be expected.
Since a hydrostatic pressure has no effect on yielding the system is equiva-
lent to a compressive stress, a, normal to the plane of the sheet metal and
the relation between a and the thickness strain, 8 3 , produces the equivalent
stress-equivalent strain characteristic curve.
A biaxial test extensometer has been described by Duncan and Johnson 5 1
and the automatic recording of equivalent stress-equivalent strain charac-
teristic curves for sheet metals was reported by Bell, Duncan and Johnson 5 2 •
Biaxial tension tests on anisotropic metals were apparently first performed
by Jackson, Smith and Lankford 5 3 . These workers assumed that the loading
was along the anisotropic axes of the metal, although they pointed out
that this assumption was incorrect. At the pole of the deformed blank,
the applied forces in the plane of the sheet act in all directions rather than
just in the x and y directions. Bramley and Mellor 54 therefore suggested
that it is preferable to average the planar properties of the metal thus imposing
rotational symmetry about the z axis. As a basis for calculation, the R values
were determined from tensile test specimens cut from sheet every 10 degrees
to the direction of rolling. The average R value. R, was determined from
the area under the curve obtained by plotting R value against orientation
to the direction of rolling. The stress-strain curve corresponding to the
value of R was also derived from the experimental data.
164 ENGINEERING PLASTICITY

For a material subjected to biaxial stress such that rotational symmetry


about the z axis exists
N=F+2H=G+2H
and R=H/G=H/F
Equation (5.81) reduces to
ii = (3/2)112[ {a;+ (J; + R((Jx- (Jyf} /(2 + R)Jll2 (6.106)
and equation (5.85) becomes
de= (2/3) 112 [ {(2 + R)/(1 + 2Rf} {(dey- Rdezf + (dex- Rdezf
+ R(dex- deyf} Jl 12 (6.107)
The normal stress, (Jz, is then related to the average tensile stress, (Jav' by
equation (6.106)
ii = (3/2) 112 {2/(2 + R) p1 2 (Jz = (3/2) 112 {(1 + R)/(2 + R)} 112 (Jav
or (Jz = {(1 + R)/2} 112 (Jav (6.108)
The accompanying strain ratios can be derived from equations (5.80) and
(6.107). Since there is no change in the strain ratios during the straining
process, the total strains can be substituted for incremental strains and
from equations (5.80)
ex : ey : ez = {(R + 1)(Jx- R(Jr- (Jz} : {(R + 1)(Jy- R(Jx- (Jz} : 2(Jz- (Jx- (JY
(6.109)
In uniaxial tension

Therefore
ex : ey : ez = R + 1 : - R : - 1
At the pole of the deformed circular blank
(6.110)
Hence, relating the average longitudinal strain, ex = eav from the experi-
mental data to the polar thickness strain ez by using equation (6.107)
8 = (2/3) 112{(2 + R)/2} 1/2ez = (2/3) 112{(2 + R)/(1 + R)} 112 eav
or ez = {2/(1 + R) }112 eav (6.111)
To convert the average tensile curve to the stress-strain curve from the
hydrostatic bulge test
(Jz = {(1 + R)/2} 1/2(Jav (6.112)
and ez = {2/(1 + R) } 112 eav (6.113)
Similar experiments have been performed on titanium and zinc by Bramley
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 165
and Mellor 55 and reasonable correlation was obtained for the titanium
but not for the zinc. The behaviour of steel and aluminium blanks has been
studied by Pearce 56 . The R values for these metals were less than unity and
Pearce reported that equation (6.112) and (6.113) could not be correlated
with his experimental results. The hydrostatic bulge test has also been used
to investigate strain-r&te effects on the characteristics of steel and aluminium
sheet metal 57 •
A general approach to the instability of sheet metal subjected to biaxial
tension has been proposed by Marciniak and Kuczyinski 58 who associate
instability with prior inhomogeneities in the sheet metal. This theory has
been extended by Sowerby and Duncan 59 and experimental results have
been reported by Venter, Johnson and deMalherbe 60 •

6.7.4 Determination of the equivalent stress-strain characteristic curve


for a sheet metal by means of the hydrostatic bulge test
If a hydrostatic bulge test is performed on a circular blank of sheet metal
the following measurements can be made:

(1) The initial thickness of the blank t 0 •


(2) The hydrostatic pressure p.
(3) The radius of curvature, p, at the pole using, for example, a spherometer
consisting of two fixed legs and a central leg operating a dial gauge. Let
d be the vertical distance of a point on the deformed surface of the blank
in the vicinity of the pole and r the horizontal distance from the pole to
the point. The radius of a circle passing through the pole and the given point
and having its centre on the vertical axis of symmetry is then given by
(6.114)
(4) The circumferential strain at the pole by, for example, marking
concentric circles on the initial blank at convenient radial intervals. The
radial movement of a point with increasing fluid pressure can be measured
with a travelling microscope. If D0 is the initial diameter and D the current
diameter of a circle, then the circumferential strain e = e1 = e2 = ln(D/D 0 )
and the equivalent strain e= 2e = In (DID 0 ). The thickness strain e3 = - 2e =
- 2ln(D/D0 ) and hence the thickness ratio t 0 /t = (D/D 0 ) 2 •

Since (j =a= pp/2t = (ppj2t 0 )(t 0 /t), the equivalent stress, a, and the
corresponding equivalent strain, e, can be determined for varying values
of fluid pressure p. The (a, e) characteristic curve for the sheet metal can
then be produced.

6. 7.5 Instability of a circular sheet metal blank in balanced biaxial


tension
It is known from hydrostatic bulge test experiments that, with ap. &nnealed
166 ENGINEERING PLASTICITY

metal, the maximum fluid pressure occurs accompanied by thinning of


the metal followed by fracture in the vicinity of the pole. Once the condition
of instability is attained fracture occurs under decreasing pressure.
The relation between the fluid pressure, p, and the polar curvature, p,
is given by equation (6.101), that is
p = 2iitjp
where t is the current thickness of the blank in the vicinity of the pole, ii =
u 1 = u 2 the membmne stresses at the pole, and u 3 the radial stress
is negligible.
At instability in balanced biaxial tension dp = 0. Therefore
dp = (2tdii 1p) + (2iidt/ p)- (2iitdp/ p 2) = o
or (dii jii) + (dt/t)- (dp/ p) = 0
and dii/ii=(dpjp)-(dt/t) (6.115)
However, - (dtjt) = de 3 , where e3 is the thickness strain= 1n(t/t0 ) =
-1n(t0 /t). Equation (6.115) can therefore be written as
(1/ii)(diijde 3 ) = 1 + (1jp)(dpjde 3 ) (6.116)
By referring to figure 6.22 and assuming spherical bulging of the blank
it will be seen that from the property of orthogonally intersecting chords
there is the geometrical relationship
a2 = h(2p- h)
Therefore (6.117)
Alternatively
2p- h = a2jh
2p = (a2jh) + h = h { 1 + (a2jh 2 )}
and p = (h/2) { 1 + (a2jh 2)} (6.118)
Also ( a2jh) + h - 2p = 0
or h2 - 2ph + a2 = 0
Therefore h = p ± (p2 _ a2)112

but h < p so that


(6.119)
Hi11 61 analysed the plastic bulging of a sheet metal blank by assuming
that the particles in the membrane describe circular paths which are ortho-
gonal to the current profile. On this assumption, the incremental circum-
ferential strain de 9 = de 1 = de 2 = dhj p. The incremental compressive
thickness strain de 3 = 2de = 2de 1 = 2dhj p. Substituting for p from equation
(6.117) produces
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 167

Integration yields
e3 = 2ln { 1 + (h 2 /a 2 )} (6.121)
Differentiating equation (6.117) gives
dpjdh = (h 2 - a2 )/2h 2
or (6.122)
Let y = { 1 + (h 2 ja 2 )} = 2phja 2 then equation (6.121) becomes
e3 = 2ln y
and de 3 jdy = 2/y = 2a 2 j(a 2 + h2 ) = a2 jph (6.123)
Also dyjdh = 2hja 2 (6.124)
Hence by combining equations (6.122), (6.123) and (6.124)
de 3 jdp = (de 3 jdy )( dyjdh )(dhjdp)
= { 2a 2 j(a 2 + h2 ) }(2h/a 2 ){ 2h 2 /(h 2 - a 2 )}
= 4h 2 /{p(h 2 - a 2 )}

Therefore
dpjde 3 = p(h 2 - a2 )j4h 2 (6.125)
and equation (6.116) becomes
( 1/0')(dO' /de 3 ) = 1 + { (h 2 - a 2 )/4h 2 }
= 1 + { (h2 - ph) j2h 2 }
since a2 = 2ph - h2 . Therefore
( 1/0')( d0'/de 3 ) = 1 + ( 1/2)- (p/2h) = ( 3/2)- (pj2h) (6.126)
From equation (6.117) pj2h = (a 2 j4h 2 ) { 1 + (h2 ja 2 )} and from
equation (6.121) exp(e 3 /2) = 1 + (h 2 ja 2 )
or h 2 /a 2 = exp(e 3 /2)- 1
and a 2 /h 2 = 1/ {exp(e 3 /2) -1}

Therefore
pj2h = exp(e 3 /2)/4{ exp(e 3 /2)- 1} (6.127)
= { 1 + (e3 /2) + .... }/4 { 1 + (e 3 /2) + (eV8) + ... - 1}
= { 1 + (e 3 /2) + .... }/4 {(e 3 /2) + (e;;s) + .... }
Rj{1 + (e3 /2)} /( 4e 3 /2) {1 + (e3 /4)}
Rj{1 +(e 3 /2)}/(4e 3 /2){1-(e 3 /4)}
168 ENGINEERING PLASTICITY

~(1/4){(2/8 3 )-1}{1-(8 3 /4)}


~ ( 1/4){ (2/8 3)- (1/2) + 1}
or pj2h ~ (1/28 3) + (1/8)
Equation (6.126) then becomes
(1/0')(d0'/d8 3) ~ 1 + (1/2)- (1/28 3)- (1/8)
~ ( 11/8)- ( 1/283) (6.128)
Assume the blank material to be strain-hardening and the equivalent
stress-equivalent strain (a, e) curve to be given approximately by
a= C(D + e)q = C(D + 83)q for balanced biaxial tension
then

Therefore
(1/a)(da/d8 3) = q/(D + 83) (6.129)
Combining equations (6.128) and (6.129)
qj(D + 83) ~ (11/8)- (1/28 3)
or
Mter rearranging
118~ + 83( - 8q + llD - 4) - 4D ~ 0
If the material is fully annealed then D = 0 and
118~ + 83( - 8q - 4) ~ 0
or
Therefore
83 ~ (8q + 4)/11 ~ (4/11)(2q + 1) (6.130)
Equation (6.130) gives the instability thickness strain in balanced biaxial
tension and shows that in a hydrostatic bulge test of a circular sheet metal
blank, no matter to what extent the material is initially prestrained the
instability strain is always greater than 4/11. This suggests that the hydrostatic
bulge test is a more suitable method than the uniaxial tensile test for determin-
ing the stress-strain relationship for a sheet metal.

6.8 HOMOGENEOUS DEFORMATION


Homogeneous deformation implies that within an element considered the
state of strain is constant throughout the element. Straight lines remain
straight after straining and parallel lines remain parallel. During homo-
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 169
geneous deformation, it follows that there is no internal shear distortion or
redundant deformation and frictional stresses are absent. During the uniaxial
tensile test described in section 6.2.1 the deformation is a homogeneous
extension until the axial force attains the maximum value, that is, correspond-
ing to the onset of instability in uniaxial tension when the extension becomes
local in the formation of a neck and the deformation is inhomogeneous.

6.8.1 Homogeneous compression of a cylindrical billet


If an isotropic cylindrical billet, initially stress free, is compressed between
perfectly smooth, rigid, parallel platens so that every element is plastically
deformed to the same extent the compression is homogeneous. The height
of the billet is reduced whilst the diameter is increased. As the compression
proceeds every transverse cross-section remains circular and the rate of
increase in diameter is the same for every transverse section. No barrelling
of the cylindrical surface occurs and therefore no redundant deformation
is involved. To ensure the absence of buckling it is assumed that the initial
height to diameter ratio of the billet is, say, less than 2.
Let the current height of billet be h, and assume that it is compressed a
small amount - dh. The axial natural strain increment, dez, is
dez =- dh/h (6.131)
The total compressive axial strain effected in reducing the height of the
billet from an initial height, H 0 , to a fmal height, H 1 , is

ez = f
H,

Ho
(- dh/h) = ln(H0 /H d (6.132)

Since the billet is isotropic, and assuming the material to be incompressible,


the radial and circumferential strains are equal.
Therefore
(6.133)
and these strains are tensile.

6.8.2 Energy dissipated during homogeneous compression of a


cylindrical billet
When the billet is compressed so that the axial height is reduced the small
amount, - dh, energy is dissipated in producing plastic deformation. If
F is the axial force exerted by the platens at any instant, the energy dissipated
is
dE= F(- dh) = azA(- dh)
where az and A are the true axial stress and current cross-sectional area of
the billet at any instant.
170 ENGINEERING PLASTICITY

The total energy dissipated in compressing the billet from an initial height,
H 0 , to a final height, H 1 , is then given by

E= f Ht F(-dh)= fHt uzA(-dh)


Ho Ho

If the billet material is incompressible, the volume ofthe billet, V, remains


constant. Thus, V = Ah or A = V /h. The energy dissipated per unit volume
is therefore

E/V = fHt
Ho
uz(- dh/h)

However, - dhjh is the compressive incremental strain dez

Therefore E/V = J: uzdez (6.134)

In general, the integral f~ ude is represented to some scale by the area


under the true stress-natural strain (u,e) characteristic curve for the billet
material for the range of natural strain 0 to e.

(a) Rigid-perfectly plastic material


Assuming the billet material to be rigid-perfectly plastic for which the
uniaxial yield stress is Y then

E/V = f•z
0
uzdez = Y
IHt (- dh/h)
Ho

(6.135)
(b) Strain-hardening material
If the material is strain-hardening and the true stress-natural strain
characteristic curve for the material in compression is approximately given
by u = Aen then

E/V= J:zuzdez=A I>~dez=Ae~+ 1 /(n+1) (6.136)

Similarly if u = C(D + e)q


E/V = {C/(q + 1)}{(D + ez)q+l -Dq+l} (6.137)
and if u = F + G{ 1 - exp ( - re)}
E/V = (F + G)ez- (G/r){ 1- exp(- rez)} (6.138)
Since no energy is dissipated in overcoming frictional resistance at the
PHENOMENOLOGICAL NATURE OF ENGINEERING METALS 171
billet-platen interfaces and there is no redundant deformation, the change
in shape of the billet effected during the compression is achieved with maxi-
mum efficiency. Homogeneous deformation is the most efficient method
of producing a change of shape of metal by plastic deformation. It is therefore
a criterion by means of which the efficiency of a metal forming process for
effecting the same change of shape can be assessed.

6.8.3 Increase in temperature during adiabatic compression


of a cylindrical billet
During compression of a cylindrical billet, the energy dissipated in producing
plastic deformation is transformed into thermal energy which results in an
increase in temperature of the billet. An estimate of the maximum increase
in temperature can be obtained by assuming the compression to be adiabatic.
In this case there is no heat transfer from the billet and all the plastic work
is converted into thermal energy.
Let p be the density of the billet material, c the specific heat, J the mechani-
cal equivalent of heat, and A.() the uniform increase in temperature for unit
mass of billet material, then
E/V = J p ell()
or A.()= (E/V)/(1 pc) (6.139)

REFERENCES
1. Bridgman, P. W., Studies in Large Plastic Flow and Fracture with
Special Emphasis on the Effects of Hydrostatic Pressure, McGraw-Hill,
New York (1952)
2. Nadai, A., Theory of Flow and Fracture of Solids, vol. 1., McGraw-Hill,
New York (1950)
3. Unksov, E. P., An Engineering Theory of Plasticity, Butterworths,
London (1961)
4. Shelton, A., On the ratio of transverse to axial strain and other tensile
properties of a cold rolled steel alloy, J. mech. Engng Sci., 3, 89 (1961)
5. Crossland, B., The effect of fluid pressure on the shear properties of
metals, Proc. Instn mech. Engrs, 169, 935 (1954)
6. Nadai, A., Theory of Flow and Fracture of Solids, vol. 2, McGraw-Hill,
New York (1963)
7. B.S. 4A4 Part I, Section 3, Aerospace series specification for test pieces
and test methods for metallic materials, part I, Tensile tests (October
1967)
8. Farren, W. S. and Taylor, G. I., The heat developed during plastic
extension of metals, Proc. R. Soc., Ser. A, 107,422 (1925)
9. Ludwik, P., Elemente der technologischen Mechanik, Springer, Berlin
(1909)
172 ENGINEERING PLASTICITY

10. Voce, E., The relationship between stress and strain for homogeneous
deformation, J. lnst. Metals, 74, 537 (1948)
11. Ramberg, W. and Osgood, W. R., Description of stress-strain curves by
three parameters, NACA Technical Note No. 902 (July, 1943)
12. Bridgman, P. W., Studies in Large Plastic Flow and Fracture with
Special Emphasis on the Effects of Hydrostatic Pressure, pp. 9-37,
38-86 and 181, McGraw-Hill, New York (1952)
13. Loizou, N. and Sims, R. B., The yield stress of pure lead in compression,
J. Mech. Phys. Solids., 1, 234 (1953)
14. Cook, M. and Larke, E. C., Resistance of copper and copper alloys
to homogeneous deformation in compression, J. Inst. Metals, 71,
371 (1945)
15. Nadai, A., Plasticity, McGraw-Hill, New York (1931)
16. Watts, A. B. and Ford, H., An experimental investigation of the yielding
of strip between smooth dies, Pro c. Instn mech. Engrs (B), 1B, 448 (1952)
17. Parker, E. R., Davis, H. E. and Flanigan, A. E., A study of the tension
test, Proc. Am. Soc. Test Mater., 46, 1159 (1946)
18. Tipper, C. F., The brittle fracture of metals at atmospheric and sub-zero
temperatures, Metal!. Rev., Inst. Metals, 2, 195 (1957)
19. Magnusson, A. W. and Baldwin, W. M., Low temperature brittleness,
J. Mech. Phys. Solids, 5, 172 (1957)
20. Slater, R. A. C. and Johnson, W., The effects of temperature, speed and
strain-rate on the force and energy required in blanking, Int. J. mech.
Sci., 9, 271 (1967)
21. Cook, P. M., True stress-strain curves for steel compression at high
temperatures and strain-rates, Proc. Conf. Prop. at High Rates of Strain,
pap. no. 2, Instn mech. Engrs., (1957)
22. Orowan, E., The cam plastometer, BISRA Report MW/F/22/50
23. Bell, J. F., The Physics of Large Deformation of Crystalline Solids,
Springer-Verlag, New York (1968)
24. Coffin, L. F. and Conrad, H., The Cryogenic Properties of Metals in
High Strength Materials, J. Wiley, New York (1965)
25. Mahtab, F. U., Johnson, W. and Slater, R. A. C., Dynamic indentation
of copper and an aluminium alloy with a conical projectile at elevated
temperatures, Proc. Instnmech. Engrs, 180,285 (1966)
26. Slater, R. A. C., Aku, S. Y. and Johnson, W., Strain rate and tempera-
ture effects during the fast upsetting of short circular cylinders of 0.55
per cent plain carbon steel at elevated temperatures, Annals of C.I.R.P.,
XXIV, 513 (1971)
27. Johnson, W. and Mellor, P. B., Engineering Plasticity, ch. 1, p. 30,
Van Nostrand Reinhold Company, London (1973)
28. Alder, J. F. and Phillips, K. A., The effect of strain-rate and temperature
on the resistance of aluminium, copper and steel to compression, J. Inst.
Metals, 83, 80 (1954)
29. Manjoine, M. J. and Nadai, A., High speed tension tests at elevated
temperatures, Proc. Am. Soc. Test Mater., 40, 822 (1940)
PHENOMENOWGICAL NATURE OF ENGINEERING METALS 173
30. Manjoine, M. J., Influence of rate of strain and temperature on yield
stresses of mild steel, J. appl. Mech. A-211 (1944)
31. Samanta, S. K., Resistance to dynamic compression of low carbon
steel and alloy steels at elevated temperatures and at high strain-rates,
Int. J. mech. Sci., 10, 613 (1968)
32. Samanta, S. K., On relating the flow stress of aluminium and copper to
strain, strain-rate and temperature, Int. J. mech. Sci., 11, 433 (1969)
33. Malvern, L. E., Experimental studies of strain rate effects and plastic
wave propagation in annealed aluminium, Proc. Amer. Soc. mech.
Engrs., p. 81, Coll. Behaviour of Materials under Dynamic Loading,
Chicago (1965)
34. Dean, T. A. and Sturgess, C. E. N., Stress-strain characteristic of
various steels over a wide range of strain-rates and temperatures,
Proc. Instn mech. Engrs, 187, (40/73), 523 (1973)
35. Hopkinson, B., Collected Scientific Papers, Cambridge University
Press (1921)
36. Campbell, J. D. and Duby, J., The yield behaviour of mild steel in
dynamic compression, Proc. R. Soc., Ser. A, 236,24 (1956)
37. Campbell, J.D. and Duby, J., Delayed yield and other dynamic loading
phenomena in a medium carbon steel. Proc. Conf on the Properties of
Materials at High Rates of Strain, p. 214, Instn. mech. Engrs (1957)
38. Goldsmith, W., Impact, Edward Arnold, London (1960)
39. Macgregor, C. W. and Fisher, J. C., A velocity-modified temperature
for plastic flow of metals J. appl. Mech., Trans. Am. Soc. mech. Engrs
(1946)
40. Inouye, K., Studies on the hot-working strength of steels (in Japanese),
Tetsu-to-Hagane, 41, 593 (1955)
41. Johnson, W., Impact Strength of Materials, p. 218, Edward Arnold,
London (1972)
42. Ripperberger, E. A., Experimental studies of strain-rate effects and
plastic wave propagation in annealed aluminium, Proc. Am. Soc. mech.
Engrs, p. 62, Coll. Behaviour of Materials under Dynamic Loading,
Chicago (1965)
43. Thomsen, E. G., Yang, C. T. and Kobayashi, S., Mechanics of Plastic
Deformation in Metal Processing, ch. 7, p. 107, Collier-Macmillan,
London (1965)
44. Atkinson, M., Assessing normal anisotropy of sheet metals, Sh. Metal
Ind. (March 1967)
45. Whiteley, R. L. and Wise, D. E., Relationship among texture, hot mill
practice and deep drawability of sheet steel, Proc. 4th Tech. Conf on
Flat Rolled Products, vol. 3, p. 47, Wiley Interscience, Chicago (1962)
46. Whiteley, R. L., How crystallographic texture controls drawability,
Metal Prog., 94 (5), 81 (November 1968)
47. Burns, R. S. and Heyer, R. H., Orientation and anisotropy in low carbon
steel sheets, Sh. Metal Ind., 35 (372), 261 (Aprill958)
48. Bauschinger, J., Ueber die Veriinderung der Elasticitiitsgrenze und die
174 ENGINEERING PLASTICITY

Festigheit des Eisens und Stah1s, Mitt. a.d. Mech. Tech. Lab., Munchen,
No. xiii (1886)
49. McLean, D., Mechanical Properties of Metals, Wiley, New York (1962)
50. Goodier, J. N. and Hodge, P. G., Elasticity and Plasticity, Wiley, New
York (1958)
51. Duncan, J. L. and Johnson, W., The use of a biaxial test extensometer,
Sh. Metal Ind., 271 (1965)
52. Bell, R., Duncan, J. L. and Johnson, W., The evolution of a prototype
machine for automatically recording the true stress-strain curve for
sheet metal using the hydrostatic bulge test, Advances in Machine
Tool Design and Research, Pergamon Press, Oxford (1965), p. 411
53. Jackson, L. R., Smith, K. F. and Lankford, W. T., Plastic flow in aniso-
tropic sheet steel, Metals Technology Tech. Pub. 2440 (1948) and J.
Metals, 1, 323 (1949)
54. Bramley, A. N. and Mellor, P. B., Plastic flow in stabilised sheet steel,
Int. J. mech. Sci., s; 101 (1966)
55. Bramley, A. N. and Mellor, P. B., Plastic anisotropy of titanium and
zinc sheet-1, Macroscopic approach, Int. J. mech. Sci., 10,211 (1966)
56. Pearce, R., Some aspects of anisotropic plasticity in sheet metals,
Int. J. mech. Sci., 10, 995 (1968)
57. Bramley, A. N. and Mellor, P. B., The effect of strain rate on the plastic
flow characteristics of steel and aluminium sheet, J. Strain Analysis,
1, 439 (1966)
58. Marciniak, Z. and Kuczynski, K., Limit strains in the processes of
stretch forming sheet metal, Int. J. mech. Sci., 9, 609 (1967)
59. Sowerby, R. and Duncan, J. L., Failure in sheet metal in biaxial tension,
Int. J. mech. Sci., 13,217 (1971)
60. Venter, R., Johnson, W. and de Malherbe, M. C., The limit strains in
inhomogeneous sheet metal in biaxial tension, Int. J. mech. Sci., 13,
299 (1971)
61. Hill, R., The theory of the plastic bulging of a metal diaphragm by
lateral pressure, Phil. Mag., Ser. 7, 41, 1133 (1950)
7
Plane Strain Plastic Deformation

7.1 BASIC CONCEPTS AND ASSUMPTIONS


In previous chapters, plastic deformation of an element of material on a
microscopic scale was considered as though the element were independent
of the body of material. For many metal forming processes such as forging,
drawing and rolling the geometrical configuration of the tooling, frictional
conditions at the workpiece-tooling interfaces and the constraint of neigh-
bouring material will result in the plastically deforming region being
subjected to afield of stress and strain which may differ from point to point.
This field of stress-strain states makes it extremely difficult to obtain a
solution to a problem. Only by making certain assumptions is it possible
to derive good first approximations to the force, energy and power required
to perform a metal forming operation and also provide useful information
about the manner in which the workpiece material deforms.
Except for small elastic zones, large unrestricted plastic flow occurs
during many of the metal forming processes. In such cases, it is reasonable
to neglect the elastic strains which is equivalent to assuming that the work-
piece material has an infinite Young's modulus of elasticity. It is also often
adequate to neglect the variation in distortion which may occur over the
deforming cross-section and assume that originally plane transverse sections
remain plane after plastic deformation. This assumption may not lead to
significant errors because it is only in processes which involve very severe
distortion, such as extrusion, that there is considerable departure in the
actual process from the assumption of plane sections remaining plane.
In developing a field-type theory to provide information about the defor-
mation from point to point in the plastically deforming region, strain-
hardening is usually neglected so that the material is assumed to flow at
constant yield stress. This hypothetical material is referred to as rigid-
perfectly plastic. Furthermore, creep and strain-rate effects are neglected.
The strain-rate at each point in a plastically deforming region is usually
different. The effect this may have on the yield stress is therefore ignored.
Inertia forces are also neglected which infers that problems are only consider-
176 ENGINEERING PLASTICITY

ed as quasi-static. During forming processes involving high values of strain,


most of the external energy utilised in producing plastic deformation is
eventually dissipated as thermal energy. Temperature gradients can therefore
arise which affect the material properties and also, for example, the lubri-
cation conditions. The thermal stresses which result from temperature
gradients are generally neglected.
At first sight, some of these assumptions may appear unrealistic. However,
since the plastic strains are generally large, it is permissible to assume an
infinite elastic modulus. Assuming the flow stress to remain constant is
closely representative of the behaviour of all engineering metals during
hot-working when the temperature is above the recrystallisation temperature
or the homologous temperature T 8 > 0.5, and also dqring the cold-working
of prestrained metals.
In addition to the above assumptions related to the workpiece material,
this chapter is devoted to a discussion of the theory which assumes that
deformation occurs under conditions such that the strain is zero in one
direction, that is, plane strain deformation. Plane strain conditions are
closely approximated during bar forging, sheet drawing and extrusion and
cold strip rolling.
The effects of introducing the various assumptions in the theory is such
that the correlation between theoretical prediction and experimental result
should not be expected to be very high. Nevertheless, the th~ory has been
proved to be very useful in many applications, despite its obvious limitations.

7.2 FUNDAMENTAL EQUATIONS FOR PLANE PLASTIC


FLOW
Plane plastic deformation implies that the displacements of elements in the
plastically deforming region all occur in parallel planes, say, planes parallel
to the (xy) plane in a system of three mutually orthogonal planes. All displace-
ments are then independent of z :
ux = uix,y), uY = uy(x,y), uz = 0 (7.1)
The wprkpiece material i~ assumed to be isotropic. In any plane
z = constant the same stress-strain relationship is applicable and the
components of stress depend only on x and y. Therefore
(7.2)
and it follows that the z direction is a principal direction. Thus, az is a
principal stress. Also
ez = Yxz = Yyz = 0 (7.3)
or in terms of the components of the plastic strain increment
dez = dyxz = dyyz = 0 (7.4)
PLANE STRAIN PLASTIC DEFORMATION 177
In addition
l;z = Yxz = Yyz = 0 (7.5)
If the material is assumed to be incompressible, then
ex+ey+ez= O
Hence (7.6)
Similarly

and (7.7)
The Levy-Mises stress-strain increment equations are
deij = a;j dA. = (3/2)(di/O') a;j
where dA. is a non-negative parameter. Therefore
deij = (aij- am)dA. where am= (ax+ ay + az)/3
Hence dex = (2/3)dA. {ax- (1/2)(ay + az)}}
dey= (2/3)dA. {ay- (1/2)(ax + az)} (7.8)
dez = (2/3)dA. {az- (1/2)(ax + ay)}
and dyxy = 't"xydA.
However, dez = 0. Therefore, from the third of equations (7.8)
az =(ax+ ay)/2 (7.9)
so that the hydrostatic stress, am, is given by
am= (ax+ ay + az)/3
= {ax+ ay + (1/2)(ax + ay)} /3 (7.10)
=(ax+ ay)/2
Hence, for plane strain deformation in the (xy) plane
(7.11)

The von Mises yield criterion in terms of the components of the stress
tensor, a;p is
(ax- ayf + (ay- az) 2 + (az- ax) 2 + 6(-r;Y + -r;z + -r;x) = 6k 2
where k is the yield stress in pure shear, but for plane strain deformation in
the (xy) plane

Therefore
178 ENGINEERING PLASTICITY

which reduces to
(1/4)(ux- uy) 2 + r;Y = P (7.12)
The principal stresses in the plastically deforming region are
0" 1.= (1/2)(ux + uy) + {(1/4)(ux- uy) 2 + r;Y} 1/2 }
0"2 = O"z = (ux + uy)/2 = p say (7.13)
u 3 = (1/2)(ux + uy)- { (1/4)(ux- uy) 2 + r;Y }1/2
The maximum shear stress in the plane of flow is given by
'!max= k = (u1- 0"3)/2
= {(1/4)(ux- uy)2 + r;y}1/2 (7.14)
Consequently, the principal stresses can be expressed as
(T 1 = P + k, (T 2 = (Tz = (Tm = p, (T 3 = P- k (7.15)
That is, the stress state at every point throughout the plastically deforming
region is characterised by the superposition of a hydrostatic stress, p, on a
pure shear stress, k, as illustrated in figure 7.1. This assumes that the stress
state for the metal forming process is predominantly compressive such that
the hydrostatic stress is compressive, that is, a hydrostatic pressure p.
The representative or equivalent stress, u, in terms of the principal stresses
is given by equation (2.46) which is
(j = (1/2)1f2{(u1- u2)2 + (u2- u3)2 + (u3- u1)2 pt2
If the von Mises yield criterion is assumed this equation reduces to
0'=(31/2/2)(u1-u3)=31i2k (7.16)
and yielding occurs when k attains the value Y/(3) 1 ' 2 , where Y is the yield
stress. However, ifthe Tresca yield criterion is applied the equivalent stress is

Figure 7.1. Stress state at a point in a plastically deforming region under plane
strain conditions characterised by a hydrostatic stress superimposed on a pure shear
stress
PLANE STRAIN PLASTIC DEFORMATION 179
ii = (}'1- (}'3 = 2k (7.17)
and yielding occurs when k attains the value Y /2. Therefore, there is no
difference between the functional relation of the stresses representing either
the von Mises or Tresca yield criteria under plane strain conditions. Conse-
quently, the yield criterion can be summarised by equation (7.12) where
k = Y/3 112 for the von Mises yield criterion and Y/2 for the Tresca yield
criterion.
The differential equations for force equilibrium, with reference to a
cartesian system of coordinates, when body forces are insignificant or
absent and inertia forces are neglected, are given by equations (2.8). For
plane strain deformation in the (xy) plane independent of z, these equations
reduce to

~:x + a;;x= 0 ) (7.18)


ar:xy+~ =0
ax oy
Equations (7.14) and (7.18) represent three equations in the three unknowns
ux, uY and r:xy· If the boundary conditions are stated only in terms of stresses,
then these equations are sufficient to define the stress distribution indepen-
dent of the strain and hence without recourse to the stress-strain increment
relations. Problems of this type are referred to as being statically determinate.
However, if displacements or velocities are specified over part of the boun-
dary, the stress-strain increment equations may have to be employed in
order to relate the stresses to the strains and the problem becomes more
complicated. Such problems are known as statically indeterminate.

7.2.1 Mohr stress circle diagram for plane strain deformation


The state of stress at some point P in a plastically deforming region may be
represented in the Mohr stress circle diagram for plane strain as given in
figure 7.2(b) and the corresponding physical plane is illustrated in
figure 7.2(a).
Because in many metal forming processes the stress states are often
predominantly compressive, it is assumed here that ux, uY and uz are all
compressive and uY > uz > ux algebraically. The Tresca yield criterion in
plane strain deformation is (}' 1 - (}' 3 = 2k, so that (}' 1 ~ (0' 2 = 0'z = 0'm) ~ (}' 3
algebraically.
The point A in the Mohr stress circle diagram corresponds to the stress
state (- ux,- -rxy) on the plane PY and the point B corresponds to the stress
state (- uY, + r:yx) on the plane PX. The sign convention used for the shear
stresses is such that r:xy or r:yx is considered positive if its sense is radially
outwards from P. The centre of the circle is C and its radius corresponds to
180 ENGINEERING PLASTICITY

O'y
+7"yx
-rxy
p
• O'x

(2)
y
(a)

(l)~ax=+k +7"
Shear stress

O'z• O'nf 0'2= IO'x+ 1/2 = P


O'x

-7"
(b)

Figure 7.2. Stress state in plane strain deformation for a rigid-perfectly plastic
material showing (a) the physical plane, and, (b) the Mohr stress circle diagram

the maximum shear stress, Tmax' that can be attained which is the yield shear
stress, k, in plane strain.
There are two mutually orthogonal planes, namely P(l) and P(2) on
which the shear stresses attain the maximum possible values Tmax = ± k,
respectively, and on which the normal stress has the value p (compressive).
The traces of these planes shown in figure 7.2(a) may be referred to as the
first and second shear lines. The stress states existing on these two planes are
represented in the Mohr stress circle diagram by the points 1 and 2, respec-
PLANE STRAIN PLASTIC DEFORMATION 181
tively. For an isotropic material, the directions of maximum shear strain-rate
coincide with the directions of maximum shear stress and these directions are
also directions of zero rate of extension or contraction.
By referring to figiire 7.2(b) it can be seen that the stress components
ux, uY and -rxy = -ryx cari be expressed in terms of the hydrostatic pressure, p,
and the yield shear stress, k, as follows:
ux = - (p + k sin 2cjJ) = - p- k sin 2cjJ (7.19)
uY = - p + k sin 2cjJ (7.20)
± 't"xy = ± k COS 2c/J (7.21)
where cjJ is the angle through which the plane PY must be rotated anticlock-
wise for it to coincide with the second shear line. It corresponds to an anti-
clockwise rotation of 24> from CA to C(2) in the Mohr stress circle diagram.
Combining equations (7.19) and (7.20) produces
Ux + UY = - 2p (7.22)
which agrees with equation (7.11) since the hydrostatic stress, p, is compres-
sive and hence negative. Also
-u1-(-um)= -u~
(7.23)
or
which is the compressive deviator stress corresponding to the principal
direction (1), and
-u3-(-um)= -u~

+Y

-'Y
Figure 7.3. Mohr circle diagram for strain-rate during the plane strain deforma-
tion of a rigid-perfectly plastic material
182 ENGINEERING PLASTICITY

or (7.24)
which is the compressive deviator stress corresponding to the principal
direction (3).

7.2.2 Mohr circle diagram for strain-rate and plastic strain increment
under plane strain conditions
Since /; 2 = de 2 /dt = 0 where t denotes time, 83 = - 81 if the material is
assumed to be incompressible such that there is no volume change, and the
equation /; 1 + 82 + i; 3 = 0 is satisfied.
The Mohr circle diagram for strain-rate is therefore as shown in figure 7.3.
It was shown in section 5.3.1. that the Mohr circle diagrams for stress and
plastic strain increment are similar and this was illustrated in figure 5.2.
The Mohr circle diagram for plastic strain increment under plane strain
conditions will therefore be similar in form to figure 7.2(b).

7.3 PLANE STRAIN SLIP LINE FIELD THEORY

7.3.1 Shear lines or slip lines


The magnitudes of the principal stresses in the plastically deforming field
are given by equations (7.13) or, alternatively, in terms of the hydrostatic
stress, p, and the yield shear stress, k, by equations (7.15). The first principal
direction is defmed as the direction of the algebraic maximum principal
stress u 1 • If () is the angle between the first principal direction and the PX
axis, as shown in figure 7.4, then the principal directions are obtained from
equations (2.22) which produce
tan 2() = 2r:xy/(ux- uy) (7.25)
Equation (7.25) gives two values of() which differ by n/2 so that the second
principal direction, which is the direction of the algebraic minimum principal
stress, u 3 , is 90° anticlockwise from the first principal direction. It should be
noted that the intermediate principal stress, u 2 , has a direction which is
normal to figure 7.4 in the direction of zero strain. The maximum shear
stresses have the values •max= ± k = ± (u 1 - u 3 )/2 and act on surfaces
which make angles of ± n/4 with the principal directions. The directions
of these surfaces on which the shear stress attains a maximum value k are
usually designated the alpha (a) and beta ({J) directions. The a direction or
first shear direction is 45° clockwise from the first principal direction and
the f3 direction or second shear direction is therefore 90° anticlockwise from
the first shear direction as illustrated in figure 7.4.
The angle which the first shear direction makes with the PX axis, when
measured anticlockwise,is 4>asindicated in figure 7.4. Therefore 4> = ()- (n/4).
Hence tan 24> = - 1/tan 2() and from equation (7.25)
(7.26)
PLANE STRAIN PLASTIC DEFORMATION 183

(I)

I
I
I

o;=-(p-k)
o;=-(p+k)

Figure 7.4. Principal stress directions and the IX and f3 directions at a point in a
plastically deforming region

It follows that
COS 2c/J = 1:xy/k (7.27)
and sin 2¢ = (ay- ax)f2k (7.28)
If curves are drawn in the (xy) plane such that at every point on each
curve the tangent coincides with a direction of maximum shear stress then
two orthogonal families of curves are obtained which are termed shear
lines or slip lines. These two families of curves are usually designated r:1. lines
and {3lines corresponding to the first and second shear lines respectively.
The stresses on a small curvilinear element bounded by slip lines are
shown in figure 7.5 and the slip lines are designated r:1. and {3 accordingly.
It should be noted that it is necessary to correctly designate the two families
of slip lines. The usual convention of achieving this is that when the r:1. and
{3lines form a right-handed coordinate system of axes then the line of action
of the algebraic maximum principal stress, a 1 , is contained in the first and
184 ENGINEERING PLASTICITY

I
Direction of the
v algebraic maximum
~lstrcssOj

I a

Figure 7.5. The hydrostatic pressure and yield shear stress on a small curvilinear
element

third quadrants. The anticlockwise rotation, c/J, of the tX line from the chosen
x direction is then considered to be positive.

7.3.2 Hencky stress equations


The state of stress at a point can be expressed by equations (7.19), (7.20) and
(7.21) in terms ofthe independent quantities p, k and c/J. The force equilibrium
equations can thus be written in terms of these quantities. From equation
(7.19): G'x + p + k sin 2c/J = 0. Differentiating with respect to x yields
oG' op oc/J
_x +- + 2k COS 2c/J- =0
ox ox ox
From the first of equations (7.18)

0(]'x = - oryx = - ~(k cos 2c/J) = 2k sin 2c/J oc/J


ox oy oy oy
Therefore
. oc/J op oc/J
2k sm 2c/J ay + ox + 2k cos 2c/J ox = 0 (7.29)

Similarly, from equation (7.20): G'Y + p- k sin c/J = 0. Differentiating with


respect to y yields
PLANE STRAIN PLASTIC DEFORMATION 185
au ap acjJ
_Y+--2kcosc/J-=0
ay ay ay
From the second of equations (7.18)
au a'l: a acjJ
= - -(k cos 2c/J) = 2k sm 2cjJ-
0

.::..::.X= - __g_
ay ax ax ax

Therefore 2k sin 2c/J acjJ + ap - 2k cos cP acjJ = 0 (7.30)


ax ay ay
Equations (7.29) and (7.30) are the partial differential equations of equili-
brium for the plane strain deformation of a rigid-perfectly plastic material
and are hyperbolic. A rigorous solution to these equations can be obtained
by the method of characteristics. The characteristic curves or characteristics
of the hyperbolic equations, in this case, coincide with the slip lines. The
concept of characteristics of partial differential equations is briefly discussed
in appendix 2. Further treatment of the subject can be found in the textbook
by Hill 1 and the monograph by Johnson, Sowerby and Haddow 2 •
However, the choice of the x and y axes is arbitrary. If PX and PY in
figure 7.4 are chosen so as to coincide with the first and second shear lines,
respectively, then the x and y axes coincide with the 0( and p directions at
the point P under consideration and cjJ = 0. The first and second shear lines
are the tangents to the 0( line and p line, respectively, at the point P.
This is equivalent to differentiating along the first and second shear lines
so that sin 2cjJ and cos 2cjJ in equations (7.29) and 7.30) have the values 0 and
1, respectively. These equations therefore reduce to the following partial
differential equations
ap + 2k acjJ = o (7.31)
ax ax

ap- 2k acjJ =o (7.32)


ay ay
Since the point P for which these equations are valid is arbitrary, it follows
that equations (7.31) and (7.32) are applicable to all points along the slip
lines. Integration thus produces the following relationships applicable to
the slip lines
p + 2kcjJ =a constant (C 1 ) along an 0( line}
(7.33)
and p - 2kcjJ = a constant (C 2 ) along a Pline
Although the constants C 1 and C 2 may be applicable to a particular
0( lin.! or p line, respectively, their values generally vary from one slip line
to another.
Equations (7.33) were first derived by Hencky 3 in 1923. It is evident that
if p and cjJ are prescribed for, say, a boundary condition then it may be
possible to proceed along constant 0( and p lines to determine the value of
186 ENGINEERING PLASTICITY

the hydrostatic pressure everywhere in the slip line field network. If the
displacements or velocities are prescribed over part of the boundary, as is
sometimes the case, then the Hencky stress equations are not sufficient to
obtain a solution. It is then necessary to use the velocity equations which are
considered in section 7.3.3.
If the yield shear stress, k, is allowed to vary, then additional integral terms
appear in the Hencky stress equations. This modification was first suggested
by Christopherson, Oxley and Palmer4 and further reported by Palmer and
Oxley 5 who required to adapt the slip line field theory for a rigid-perfectly
plastic material to be applicable to a strain-hardening material during
orthogonal machining. However, it is then not clear how the strain distri-
bution is affected.

7.3.3 Geiringer velocity equations


Consider an element in a plastically deforming region in the vicinity of the
point P shown in figure 7.6. Let its velocity at a given instant be V and the
components of this velocity along the IX and f3 slip lines be u and v, respectively.
If the element crosses a slip line it may be subjected to shear deformation
in the shear direction. However, since it is assumed that only maximum shear
stresses can exist along shear lines and the normal stresses are everywhere
equal to the hydrostatic pressure, no extension or contraction can occur
along the slip lines although the element can distort in pure shear. Therefore
8uj8sa. = 8vj8sp = 0 (7.34)
where sa. and sp denote distances measured along the IX and f3 lines,
respectively.
There may, however, be a change in velocity due to a change in direction
of shear which changes the magnitude of the velocity component normal to
the line.

Figure 7.6. Component velocities of a particle in the x andy directions and IX


and fJ directions at a point within a plastically deforming region
PLANE STRAIN PLASTIC DEFORMATION 187
Resolving the velocity components u and v along the shear lines in the
PX and PY directions gives
ux = u cos cjJ - v sin cjJ (7.35)
vY = v cos cjJ + u sin cjJ (7.36)
Differentiating equation (7.35) with respect to x produces
oux . o¢ cos¢.-- . ov)
ocjJ+ smcjJ-
ou ( vcos¢ -
-=u.- sm¢-+ (7.37)
ox ox ox ox OX
and differentiating equation (7.36) with respect to y produces
oc/J
ov =v. -sm¢-+cos¢-+
. oc/J
ov ( ucos¢-+sm ou)
. cjJ- (7.38)
_Y
oy oy oy oy oy
If the PX and PY axes are rotated through an angle cjJ anticlockwise they
coincide with the tangents to the rx and f3 slip lines, respectively, and cjJ = 0.
Therefore

- oux
ou -[ - J
-- ou- voc/J
OS~- ox 4>=0- ox
-- 0
ox-
(7.39)

and ~ = [ ovy
osp
J
oy <~>=o
= ov + u oc/J = 0
oy oy
(7.40)

or du - v d¢ = 0 along an rx line}
(7.41)
and dv + u d¢ = 0 along a f3 line
Equations (7.41) are the velocity compatibility equations which were first
derived by Geiringer 6 in 1930.
If the problem is statically determinate, the slip line field and the stresses
can be defined from equations (7.33) and the stress boundary conditions.
The velocities can then be determined from equations (7.41) using the velocity
boundary conditions. However, if the problem is statically indeterminate
when the stress boundary conditions are insufficient to obtain a unique slip
line field then the Hencky stress equations must be solved simultaneously
with the Geiringer velocity equations using both the stress boundary condi-
tions and the velocity boundary conditions. Except for the cases where the
slip line fields are of the simplest kind, for example, the single triangle or
circular sector, the numerical solution to statically indeterminate problems
is extremely difficult and must usually be performed by trial and error.

7.3.4 The hodograph


It is possible in many cases, however, to construct a graphical representation
of the velocity at each point in a plastically deforming region which is known
as a hodograph from the Greek hodus meaning 'the way'. This diagram, first
188 ENGINEERING PLASTICITY

Physical plane

0
Hodograph

Figure 7. 7. An element of a slip line field and the corresponding hodograph

named by Prager 7 permits evaluation of the magnitude and direction of


the velocity at a point which is indicated as a vector. The velocity along a
slip line maps on the hodograph as a line. Such lines are transformed slip
lines and are parallel to the components u and v of the velocity vectors. The
hodograph enables the velocity distribution associated with a slip line field
to be established without great difficulty. The strain-rate components and
the flow paths of material elements can be determined. It is then possible to
establish that a proposed slip line field is kinematically admissible and
compatible with the imposed conditions. Referring to figure 7.7, consider
two points A and B which are an infinitesimal distance apart on a slip line.
Since the extension or contraction in the direction tangential to the slip
line is zero, the velocity of B relative to A has a direction normal to AB. A
fixed origin, 0, representing zero velocity is chosen in the hodograph plane.
Bound vectors which represent the velocity at points in the slip line field
can then be drawn from this origin. A point in the slip line field consequently
maps, in the hodograph plane, as the terminal of the corresponding bound
vector. The points A and B map as the points a and b in the hodograph
plane shown in figure 7.7 where oa
= va =velocity of the point A relative to
earth and Ob = vb = velocity of the point B relative to earth.
~
The vector ab is the velocity of B relative to A, avb, which must have a
direction normal to AB, and the element AB of the slip line maps as the
element ab in the hodograph which is normal to AB. It follows that the
vectors in the hodograph form an orthogonal network having the same
geometrical properties as the slip line field. A non-deforming region moving
in translation maps as a point in the hodograph plane.
PLANE STRAIN PLASTIC D,EFORMA TION 189
7.3.5 Geometrical properties of slip lines
A plastically deforming region ABCD shown in figure 7.8 is bounded by
two r:t. lines, AB and DC, and two f3 lines, AD and BC. It is assumed that the
x direction is as shown and an anticlockwise rotation from the x direction
defines a positive value of the angle¢.
The difference in the hydrostatic pressure between C and A can be derived
in two possible ways using the Hencky stress equations as follows :

(1) A --+ B along r:t.line, p + 2k¢ = C1


Therefore (7.42)
B --+ C along f3 line, p - 2k¢ = C 2
Therefore Ps - 2k¢ 8 = Pc - 2krpc (7.43)
Subtracting equation (7.42) from equation (7.43) produces
- 4k¢8 = Pc ~ PA- 2k¢A- 2k¢c

and the difference in hydrostatic pressure between C ang A is


Pc- PA = 2k(¢A + ¢c- 2¢s) (7.44)

(2) A --+ D along f3 line, p - 2k¢ = C 3


Therefore (7.45)
D--+ C along r:t.line, p + 2k¢ = C 4

Figure 7.8. A plastically deforming region ABCD bounded by two 11. lines and
two {3 lines demonstrating Hencky's first theorem
190 ENGINEERING PLASTICITY

Therefore Po + 2k<f>o = Pc + 2k<f>c (7.46)


Subtracting equation (7.45) from equation (7.46) produces
4k</> 0 = Pc- PA + 2k<f>c + 2k</>A
or Pc- PA = - 2k(</>A + <f>c- 2</>o) (7.47)
Combining equations (7.44) and 7.47) then gives
</>A- <Po= <f>a- <f>c (7.48)

Equation (7.48) implies that there is a constant angle between the tangents
to two slip lines of one family (a) at their intersection with a slip line of the
other family (/3). This is known as H encky' s first theorem.
This theorem is of considerable importance in the numerical and graphical
construction of slip line fields. A useful deduction from this theorem is
that if a segment of a slip line, cut off by two slip lines of the other family,
is a straight line then it follows that all the other segments cut off by the same
two slip lines of the other family will also be straight lines. The straight
segments are the common normals of the intersecting slip lines of the other
family. These slip lines thus have a common evolute and the straight segments
are all of the same length.
Consider a pair of a and f3 slip lines and let their radii of curvature be + R
and + S, respectively, when their curvatures are as shown in figure 7.9.
The length of the elemental segment PQ of the a line
t5sa = Rt5</> (7.49)
where sa is the distance measured along the a line.

Figure 7.9. Convention for the radius of curvature of a slip line


PLANE STRAIN PLASTIC DEFORMATION 191
f3 f3

Figure 7.10. An infinitesimal curvilinear element bounded by two ex lines and


two fJ lines demonstrating Hencky's second theorem

The average curvature of the segment PQ is ()<jJj()sa. Similarly, the length


of the elemental segment LM of the {3 line
(7.50)
where sp is the distance measured along the {3 line.
The average curvature of the segment LM is b</J/bsp· As ()sa and bsp
approach zero the actual curvatures at the points P and L on the rx line and
{3 line, respectively, are given by
d</Jjdsa = 1/R } (7_51 )
and d¢/dsp = - 1/S
In figure 7.10, ABCD is an infinitesimal curvilinear element bounded by
two a lines which are an infinitely small distance bsp apart and two {3 lines
which are an infinitely small distance bsa apart.
AD= bsp = - S()<jJ
where bsp is the length of the segment on the {3 line cut off by the two rx lines.
d
BC=bsp+-d (bsp)bsa= -b</J(S-bsa)= -Sb</J+b</Jbsa
sa
d
Therefore BC- AD= -d (bsp)()sa = bsab</J
sa
d
or -d (bsp) = b</J
sa
but bsp= -Sb</J
d
Hence - ( - S()<jJ) = J¢
dsa
However, from Hencky's first theorem, the angle (j <P between the tangents
192 ENGINEERING PLASTICITY

of any two slip lines of one family at their intersection with a slip line of the
other family remains constant.
Therefore dS/ds(l = -1} (7.52)
and dR/dsp = -1
The second of equations (7.52) can be derived in a manner similar to the first.
From equations (7.52) it follows that
dS= -ds(l= -Rd¢
and dR = - dsp = + Sd¢
Therefore
dS + Rd¢ = 0 along an IX line }
(7.53)
dR - Sd¢ = 0 along a f3line
Equations (7.52) or, alternatively, equations (7.53) constitute Hencky's
second theorem which states that on moving along a slip line, the radii of
curvature of the slip lines of the other family at the points of intersection
change by an amount equal to the distance traversed.
If the distance traversed along a slip line is far enough in the appropriate
direction and the plastically deforming region extends far enough, then the
radii of curvature of the intersecting slip lines become zero. Also, on moving
along a slip line, the centres of curvature of the intersecting slip lines form
an involute of the slip line.
Other restrictions are placed on the geometry of the slip line field if the
Hencky stress equations are to be valid. However, since these other properties
of slip lines are not frequently used in solving problems of plasticity theory
no reference will be made here. The proofs of other theorems are to be found
in the textbook by Prager and Hodge 8 .

7.3.6 Slip line fields for simple stress states


Some typical slip line field patterns for simple stress states frequently encoun-
tered in metal forming problems are those presented in figure 7.11. Figure
7.11(a) shows a slip line field generated by two orthogonalfamilies of parallel
straight lines. It follows directly from the Hencky stress equations (7.33) that
if the slip lines are straight then the angle, ¢, is constant along the slip lines
and the hydrostatic pressure, p, remains constant, the stress components
ax, aY and az must also be constant. The slip line field of figure 7.11(a) thus
represents a uniform stress state.
The slip line field shown in figure 7.11(b) comprises a set of radial straight
lines emanating from a point 0, say IX lines, and a family of concentric
circular arcs, say f3lines. From the first of Hencky's stress equations (7.33),
since cfJ is constant along an IX line, the hydrostatic pressure, p, must also be
constant along an a line. From the second ofHencky's stress equations (7.33),
PLANE STRAIN PLASTIC DEFORMATION 193

(a) (b)

(c) (d)

Figure 7.11. Some slip line fields of simple stress states

since ¢ varies linearly with distance along a fJ line, the hydrostatic pressure,
p, varies linearly with distance along a fJ line. It follows that the hydrostatic
pressure is constant in the radial direction and varies linearly with the angle
measured from the x axis. The stress components can be determined from
equations (7.19), (7.20) and (7.21). This type of slip line field is known as
the centred fan.
It should be noted that the centre of the fan, 0, is a point of stress singularity
since it can have any one of an infinite number of values. In metal forming
processes, stress singularities usually occur at die corners or at sudden
changes in cross-section of tooling and represent points where there are
rapid changes in stress and velocity.
A general simple stress state consists of one family of straight slip lines,
say oc lines, while the second family of fJ lines is generated by orthogonal
curves as shown in figure 7.11(c). For a simple stress state, the straight slip
lines are tangential to the envelope of the family as shown in figure 7.11(d).
This envelope is called the limit curve. In the case shown, the family of fJ lines
is generated by equidistant curves which are involutes with respect to the
limit curve.
From the cases considered, it can be inferred that a region adjoining a
region of uniform stress state is always in a state of simple stress. Assume
194 ENGINEERING PLASTICITY

(b)

Figure 7.12. Slip line fields comprising (a) a region of simple stress state adjoining
a region of uniform stress state, and (b) two separate regions of uniform stress state
combined with a central centre fan field

that the region A in figure 7.12(a) is one of uniform stress state. A segment of
the straight slip line, SL, which bounds the region A is one of the family of
Plines. Along this slip line the hydrostatic pressure remains constant. The
region B must be contiguous with the solution along the boundary, SL,
and therefore must be a region of simple stress state.
Regions of uniform stress state can be combined with regions of simple

Figure 7.13. A more complicated slip line field which combines three regions of
uniform stress state with two centre fan fields 1
PLANE STRAIN PLASTIC DEFORMATION 195
stress state in various ways. Figure 7.12(b) illustrates a slip line field consisting
of two separate regions of uniform stress state, A and C, combined with a
central centre fan field B. The stress field is continuous throughout the
regions A + B + C except at the centre, 0, which is a stress singularity.
A more complicated case is illustrated in figure 7.13. In this particular field
the regions A, C and E are regions of uniform stress state which are combined
with two centre fan fields B and D. Again, the stresses are continuous except
at the point 0. These examples of slip line fields are frequently used in
obtaining solutions to particular metal forming problems.

7.3. 7 Stress discontinuities


The normal stress-shear stress (a, r) plane which contains the Mohr stress
circle diagram to represent the stress state (ax, aY, rxy) at any point in, for
example, a plane strain deformation is usually known as the stress plane.
For a rigid-perfectly plastic material the radius of the Mohr stress circle
cannot exceed the pure yield shear stress, k, and if the radius is less than k
the material remains rigid and no deformation occurs.
Assume the resultant stress vector at a point P in a plane plastic flow to
bear on a plane normal to the plane of flow. Let the normal and shear stress
components be a and r, respectively. If r < k, then two Mohr stress circles
of radius, k, can be drawn through the point (a, r) in the stress plane. Two
solutions to the stress state are thus possible corresponding to the left-hand
and right-hand circles shown in figure 7.14. That corresponding to the
right-hand circle is known as the strong solution and that corresponding to
the left-hand circle is known as the weak solution. However, in most appli-
cations the appropriate solution to be chosen is evident. The point (a, r) and

Physical plan~

Strus plan~

Figure 7.14. Strong and weak solutions in the stress plane


196 ENGINEERING PLASTICITY

the poles of the Mohr stress circles for the strong and weak solutions are
collinear. The pole of a Mohr stress circle is defined as the point from which
lines parallel to the respective planes in the material intersect the Mohr stress
circle at points representing the stress state on those planes.
The existence of strong and weak solutions suggests that it is possible
for a surface ofstress discontinuity to occur in a rigid-perfectly plastic material
across which there is a large stress gradient. For an engineering metal a
surface of stress discontinuity may be considered as the limit of an elastic
layer in an elastic-plastic material which coincides with it.
If the material is deforming plastically, each of the radii of the two circles
then equals the pure yield shear stress k. A line of stress discontinuity can

Line of strns discontinuity

a slip lint: for region (I)

Physical plane

Mohr stress circle


for region (I) +'I

-(T --- 0

-r
Stress plane
Figure 7.15. A stress discontinuity represented by the Mohr stress circle diagram
PLANE STRAIN PLASTIC DEFORMATION 197
never coincide with a slip line since the two circles of radius, k, can only
intersect at the points -r = ± k when they are coincident.
At every point along a line of stress discontinuity, C, shown in the physical
plane of figure 7.15 the resultant stress vector on one side must be equal in
magnitude, have the same direction but opposite sense to that on the other
side. For the element considered in a plane plastic flow this infers that the
compressive normal stress, un, and the shear stress, -r, are continuous across
the line of discontinuity but the compressive tangential stress, ul' may be
discontinuous. The stress plane of figure 7.15 illustrates the Mohr stress
circle diagrams for the two states of stress existing in the regions (1) and (2)
on the two sides of the line of stress discontinuity.
It follows from the yield criterion that if the material on each side of the
line of stress discontinuity is plastically deforming
(}'t = (}'n+- 2(k2- 't2)1/2 (7.54)
Hence
(}'~1) = (J'n _ 2(k2 _ 't2)1/2 (7.55)
which corresponds to the strong solution, and
u:2) = O'n + 2(k2 - 't2)1/2 (7.56)
corresponding to the weak solution.
It should be noted that the resultant compressive stress vector in the
regions (1) and (2) on each side of the line of stress discontinuity is represented
by OA. Also P 1 is the pole for region (1), P 2 is the pole for region (2), P 1I 1 is
the tX direction for region (1), P 2 12 is the tX direction for region (2).
The hydrostatic pressures are
p 1 = (un + up>)/2 for region (1) } (7.57)
P2 = (un + u:Z>)/2 for region (2)
It follows that the sudden change in hydrostatic pressure across the line
of discontinuity when changing from a strong solution to a weak solution
is given by
(7.58)
indicating a decrease in the magnitude of the hydrostatic pressure. From
the stress plane of figure 7.15, LI 1 P 1 A= L 12 P 2 B=cf>.
The tX lines on each side of the line of stress discontinuity therefore make
an angle cf> with the line P 1 AP 2 which is parallel to the tangent to the line
of stress discontim,1ity.
At the line of stress discontinuity, the slip lines are thus reflected as
illustrated in the physical plane of figure 7.15 where it can be seen that the
radius of each Mohr stress circle must be rotated through the same angle,
2cf>, to coincide with the maximum shear stress, k, where cf> is the angle
between the slip line and the line of stress discontinuity.
198 ENGINEERING PLASTICITY

7.3.8 Velocity discontinuities


Because a rigid-perfectly plastic material is assumed, the displacements
throughout the deforming material need not be continuous. It is thus possible
for there to be relative slipping between neighbouring zones in the deforming
material. Suppose that along some line, L, as illustrated in figure 7.16(a),
the stresses are continuous but the velocity vector is discontinuous. At any
point P on the line, let the material on one side of the line, L, in the region (1)
have a velocity V1 with components v1 and u 1 , normal and tangential to
the line L, respectively, and on the other side in the region (2) its velocity is
changed to V2 with corresponding components v2 and u2 • If the material is
neither to 'pile-up' nor form a void at the point P, then the normal compo-
nents of velocity must be equal, that is v1 = v2 . However, no such restriction
is imposed on the tangential components. It is then possible for there to
exist a velocity discontinuity of magnitude, V* = u 1 - u2 , tangential to the
line L at the point P as shown in the hodograph of figure 7.16(b).
The line of velocity discontinuity, L, is the limit position of the layer in
which the normal velocity components are constant whilst the tangential
velocity component changes rapidly through the thickness of the layer. As

Tangent to the line of velocity


discontinuity L at point P

(a) Physical plane

(b) Hodograph

Figure 7.16. A velocity discontinuity and its representation in the hodograph


PLANE STRAIN PLASTIC DEFORMATION 199
the thickness of the layer diminishes the rate of shear strain becomes infmite.
This implies that, in the limit, the direction of the line of velocity discontinuity
must coincide with the direction of the slip line. A line of velocity discontinuity
must therefore be a slip line or the envelope of slip lines and the appropriate
one of the Geiringer equations (7.41) must be satisfied on both sides of the
discontinuity. The velocity component, u, can be discontinuous along an oc
line and the component, v, along a f3line. During the plastic deformation of an
engineering metal, when flowing from an elastic region to a plastically
deforming region, a line of velocity discontinuity will be replaced by a fmite
layer or transition region in which the shear strain rate is very high.
From equations (7.41)

u= I
v dt/J + constant, along an oc line

v= I
u dt/J +constant, along a {3line

Since the velocity component, v, is continuous along an oc line and u on a


f3 line, the sudden change or jump in u or v is constant along lines of velocity
discontinuity corresponding to oc and f3lines, respectively.
The magnitude of the tangential stress along a line of velocity discontinuity
is equal to ± k. In passing through such a line, an element experiences a fmite
shear in the direction in which the tangential stresses act and thus changes
its direction of motion. The jump in velocity, for example, V* = u 1 - u2
along an oc line and the sense of the tangential stress, -r, are related by the
condition that the plastic energy dissipation be positive: -r(u 1 - u 2 ) > 0.
Therefore, if the jump V* = u 1 - u 2 > 0, -r = + k and if V* = u 1 - u 2 < 0
then -r= -k.
A general theory of the discontinuities which may occur in a plastic
medium is discussed in the textbook by Thomas 9 .

7.3.9 Stress boundary conditions


Certain limitations are imposed on the angles at which slip lines can intersect
boundaries in order that the yield criterion is not violated. The following
are examples of various boundary conditions which are most frequently
encountered in slip line field theory.

(a) Stress free surface


A plastically deforming region sometimes extends to the free surface of
the workpiece beyond the confines of the tooling. This occurs, for example,
in the vicinity of a flat-faced punch indenting a semi-infinite mass shown in
figure 7.17(a).
At the stress free surface there are no normal or shear components of
stress. The stress free surface is, therefore, a principal plane on which the
200 ENGINEERING PLASTICITY

Direction of algebraic
maximum principal stress

fJ
a
Ia I

+k +'1"

-'1"
(b)

Figure 7.17. (a) Intersection of the slip lines at a point Pin a stress free surface;
(b) Mohr stress circle diagram indicating the principal stress state at point P

principal stress is zero. It follows that the direction tangential to the free
surface is a principal stress direction.
The slip lines indicate directions of maximum shear stress at any point
in the material and consequently intersect the free surface at angles of ± n/4.
The surface in the vicinity of the punch yields under the influence of a com-
pressive force. The normal stress at the point P in the free surface can there-
fore be considered as a zero compressive stress. Since the other principal
stresses are also compressive and have greater magnitudes, the zero normal
stress is the algebraic maximum principal stress, that is, G 1 = 0. The Tresca
yield criterion is satisfied if
(11-(13=2k
Hence (13 = - 2k
and G2 = -k=p
PLANE STRAIN PLASTIC DEFORMATION 201
Thus, at the point P in the stress free surface, the algebraic maximum
principal stress (J 1 = 0, the algebraic minimum principal stress (J 3 = 2k
(compressive) and the intermediate principal stress, which is the hydrostatic
pressure at the point P, is a 2 = k (compressive)= p. This principal stress
state at the point P in the stress free surface is verified by the Mohr stress
circle diagram of figure 7.17(b).
The algebraic maximum principal stress, (J 1 , has its direction contained
in the first and third quadrants of the right-handed rx-/3 coordinate system.
Hence the rx and f3 lines are designated as shown in figure 7.17(a). If the
surface had yielded in tension instead of compression then the rx and f3 lines
would have been interchanged.

(b) Frictionless interface


A smooth die, tool face or a constraint constitutes a frictionless interface

Smooth die,tool face or constraint


Oj;tO >2k

Frictionlus
intuface

(a)

+7"
+k

-k
(j3* 0 >2k
-7"
(b)
Figure 7.18. (a) Intersection of the slip lines at a point Pin a frictionless interface
(smooth boundary); (b) Mohr stress circle diagram indicating the principal stress state
at point P
202 ENGINEERING PLASTICITY

when in contact with the workpiece such that frictional resistance is absent.
If the interface is frictionless then there can be no resultant shear stress
tangential to the interface. The interface is consequently a principal plane
and, as for the stress free surface, the slip lines must intersect the interface
at angles of ± n/4, as shown in figure 7.18(a). However, unlike the stress
free surface the normal stress at the interface is not likely to be zero.
The normal and tangential stresses at the point P in the interface must
satisfy a yield criterion. A knowledge of the magnitudes of these stresses
then permit the slip lines to be designated. Assume the normal compressive
stress at the frictionless interface, which is therefore a principal stress, due
to the tool pressure, say, is most likely to have the maximum magnitude.
This principal stress is then the algebraic minimum principal stress
a 3 =I= 0 > 2k compressive and the algebraic maximum principal stress
a 1 < 0. Referring to the Mohr stress circle diagram shown in figure 7.18(b)
for the principal stress state at the point P in the interface it can be seen that
the intermediate principal stress, which is the hydrostatic pressure at the
point P, is given by a 2 = p =I= k compressive.
Since the direction of the algebraic maximum principal stress, that is,
a 1 is contained in the first and third quadrants of the right-handed rx-/3
coordinate system, the rx and f3 lines are as designated in figure 7.18(a).

(c) Coulomb friction present at the interface


Coulomb friction implies that the coefficient of friction, /1, is constant at
every point in the interface, that is, the ratio shear stress/normal stress is a
constant.
Let the normal compressive stress at a point P in the interface due to the
tool pressure be q, then the resultant shear stress tangential to the interface
with Coulomb friction present is rxy = 11q. The sense of this shear stress is
assumed to be that shown in figure 7.19(a) and is negative. The resultant
shear force exerted on any element ABC of the interface is balanced by the
resolved components of the forces acting on the slip lines as illustrated in
figure 7.19(b). One slip line will therefore intersect the interface at some angle
cp < n/4. If the forces exerted on the element ABC are resolved in a direction
tangential to the interface and equated to zero for equilibrium, then it will
be found that
cos2cp = Mlk (7.59)
Hence 2cp = cos- 1 (M/k)
or cp =cos- 1 (M/k)/2 (7.60)
The angle cp can also be determined from the Mohr stress circle diagram
given in figure 7.19(c). However, the solution implies a knowledge of the
normal compressive stress, q, before the field can be constructed. In fact, the
magnitude of q cannot be determined until the field is completed. An iterative
procedure is therefore required. The first approach is made by determining
PLANE STRAIN PLASTIC DEFORMATION 203

(a) (b)

+k

_,.
(c)

Figure 7.19. (a) Intersection of the slip lines at a point Pin an interface where
Coulomb friction is present (partially rough boundary); (b) stresses acting on an element
at the point P; (c) Mohr stress circle diagram for the stress state at the point P

the magnitude of q by assuming the interface to be frictionless. This is suffi-


ciently accurate in some instances. In general, q varies from point to point
along the interface. It follows that the slip lines will intersect the interface
at continuously changing angles. For the point P considered, the direction
of the plane on which the algebraic maximum principal stress, a 1 , acts is
given by </J + (n/4) anticlockwise from the interface. The IX and p lines are
thus designated as shown in figure 7.19(a). Under these conditions the
interface is usually referred to as being partially rough.
(d) Perfectly rough interface
Particularly during hot forming, the frictional stress at the metal-platen
interfaces can become so high that the workpiece material will yield in shear
at the interfaces. In this case, the frictional stress attains the maximum
possible value which is the yield shear stress of the workpiece material k,
that is, txy = ± k. In general
txy = ± k COS 2</J
204 ENGINEERING PLASTICITY

q "'"

(a)

+k +'T'

(b)

Figure 7.20. (a) Intersection of the slip lines at a point P, tangentially and
normally, to a perfectly rough interface; (b) Mohr stress circle diagram for the stress
state at the point P

Hence cos 2¢ = ±1
and 2¢ = Oand n
or ¢ = Oand n/2
It follows that one slip line meets the interface tangentially and the other
normally, as shown in figure 7.20(a). The corresponding Mohr stress circle
diagram for the stress state at the point P considered, is given in figure 7.20(b).
From this diagram it can be seen that the plane of the algebraic maximum
principal stress, u 1 , has a direction which is at an angle n/4 anticlockwise
from the interface as shown in figure 7.20(a). The a and f3lines are therefore
designated as in this figure. Under these conditions the interface is usually
referred to as being perfectly rough.

7.3.10 Prager's geometrical construction of slip line fields


A geometrical construction which recognises the fact that a hodograph
PLANE STRAIN PLASTIC DEFORMATION 205
consists of transformed slip lines and that the state of stress for plane plastic
strain at a point on a slip line traces out a cycloid has been suggested by
Prager (reference 7, pp. 1-26). This method of solution has advantages in
certain circumstances, for example when the boundary conditions are mixed.
It can also provide a better insight into the principles underlying slip line
field theory.
The construction is carried out on three planes which are the physical
plane, the stress plane and the velocity plane. In the Prager construction of
the stress plane the Mohr stress circle diagram is of importance. Conse-
quently, a review of the Mohr stress circle diagram and an extension of some
of its features are first considered.
In the physical plane of figure 7.21(a) is shown a point P which is in a
field of plane plastic strain. The normal and shear components of the resul-
tant stress, ur, acting on the plane PX passing through Pare u and -r, respec-
tively, and assumed to be kr.own. It is further assumed that the stress state

Ia) Physical plane

+k +'T

-k
Parallel to PX

-'T
lbl Stress plane

Figure 7.21. Stresses at a point on various planes represented in (a) the physical
plane, and (b) the stress plane, illustrating the pole of the Mohr stress circle diagram
206 ENGINEERING PLASTICITY

in the plastically deforming region is predominantly compressive. A shear


stress which has a sense radially outwards or will cause the element to
rotate clockwise is considered positive. The shear stress, T, on the plane PX
is therefore negative.
It is now possible to draw the Mohr stress circle, since the radius. of the
circle represents a magnitude k, which is the yield shear stress of the deforming
material. The stress state ( - CJ, - T) at the point P on the plane PX is shown
as the point A in the stress plane of figure 7.21(b). If it is now required to
determine the stress state on another plane, say PR, passing through the
point P which is inclined at an angle + o:, that is, o: anticlockwise, to the plane
PX then the radius of the circle is rotated through an angle + 2o:, that is,
anticlockwise from CA as shown in the stress plane.
Since the arc of a circle subtends twice the angle at the centre that it does
on the circumference, the pole of the Mohr stress circle may be defined as
that point from which lines drawn parallel to respective planes in the deform-
ing material intersect the Mohr stress circle at points representing the states
of stress on these planes. Thus, if the resultant stress, CJr, is known on the
plane PX, the Mohr stress circle can be drawn and a line drawn from the
point A, representing the stress state (- CJ, - -r), parallel to the plane PX
in the physical plane to intersect the Mohr stress circle thus defining the
position of the pole. To determine the stress state on any other plane PR,
rotated + o: from the plane PX, a line is then drawn from the pole parallel
to PR. Where this line intersects, the Mohr stress circle in the point B defines
the required stress state giving CJ~, CJ' and -r' and the angle y'.
The slip lines are coincident with those planes on which the shear stress
attains its maximum values ± k. By joining the pole to the top and bottom
points of the Mohr stress circle, the directions of the o: and f3lines, respectively,
are defined as shown by short chain lines in the stress plane and reproduced
in the physical plane.
In moving along a slip line, the Hencky stress equations give p ± 2k¢ = a
constant or /Jp = +2k¢. Since the abscissa of the centre, C, of the Mohr
stress circle is given by CJ = - p, it can be seen that the Mohr stress circle
moves by an amount equal to + 2k/J¢, where />¢ is the change in angular
displacement involved in moving along the slip line.
Consequently for a change of angle />¢ in the physical plane there is
a rotation of the radius of the Mohr stress circle by an amount 2/J¢ and since
the radius of the circle is± k, +2k/J¢ also represents the peripheral move-
ment of the pole. Thus, in moving along a slip line, the Mohr stress circle can
be considered to roll without slipping along either the top or bottom tangent
at a distance ± k from the normal stress, CJ, axis and the locus of the pole is
a cycloid.
This is illustrated for three arbitrary points, P 1 , P 2 , and P 3 along an
o: line which are shown in the physical plane of figure 7.22(a). In the stress
plane of figure 7.22(b) the Mohr stress circles with centres, C 10 C 2 , and C 3
and radii k, represent the stress states at the points, P 1 , P 2 , and P 3 , respec-
PLANE STRAIN PLASTIC DEFORMATION 207
-~<P a line

f3 linn

(a) Physical plane:

1---2k~<P -1 +'T

----~~~~~~~<-~~-------+-+k

Cycloid generated
by pole of Mohr p2
stress circle
rolling along top P3
tangent 'T= + k t------------------ -=------1
-'T
(b) Stress plane:

Figure 7.22. Representation of the state of stress along a slip line illustrating
(a) slip lines in the physical plane, and (b) cycloid locus of the pole, P, in the stress plane

tively. The Mohr stress circle can therefore be considered to roll without
slipping along the top tangent at a distance r = + k from the normal stress,
u, axis. Since the angle, ¢, decreases when moving along the il( line from the
point P 1 to either P 2 or P 3 the circle rolls so that the radius rotates clockwise.
If the angle ¢ had increased then the circle would roll without slipping along
the tangent r = + k so that the radius rotates anticlockwise. Points such as
P 1 , P 2 , P 3 in the slip line field are mapped in the stress plane by the pole of
the corresponding Mohr stress circle. The directions of the il( and f3 slip lines
correspond to the short chain lines drawn from the pole to the points, + k
and - k, respectively, on the appropriate Mohr stress circle. As the Mohr
stress circle rolls without slipping along the top tangent, r = + k, a cycloid
is generated by the locus of the pole.
The rules for mapping the slip line field in the stress plane can be stated as
follows : (a) to map an il( line, roll the Mohr stress circle without slipping
along the tangent r = + k; (b) to map a f3 line, roll the Mohr stress circle
208 ENGINEERING PLASTICITY

without slipping along the tangent -r = - k; (c) when the angle <P increases
the circle rolls so that a radius rotates anticlockwise, and when <P decreases
the circle rolls so that a radius rotates clockwise.
Once the slip lines have been established the velocity plane or hodograph
can be constructed from the known velocity boundary conditions using the
orthogonality relation which exists between the slip lines and the hodograph.
The above discussion is intended to present an outline of Prager's geo-
metrical construction of slip line fields. For further discussion and examples
of application the reader is referred to the original paper by Prager 7 and
·other papers by Prager 10 , Ford 11 and Alexander 12 •

7.3.11 Requirements for a complete solqpon


A complete solution to the plane straip deformation of a rigid-perfectly
plastic material requires :

(1) A statically admissible stress field which satisfies the equilibrium


equations and the stress boundary conditions and nowhere is the yield
criterion to be violated.
(2) A kinematically admissible velocity field that is compatible with the
stress field such that the appropriate stress-strain rate equations are satisfied.
(3) The rate of plastic energy dissipation to be positive everywhere that
deformation occurs.

A slip line field solution that is kinematically admissible includes a


statically admissible extension of the stress field into the non-deforming
regions adjacent to the plastic regions and where the plastic energy dissipa-
tion is positive everywhere that deformation occurs is referred to as a
complete solution.
If two or more complete solutions to a problem can be determined then
the stress fields are unique except, perhaps, in the common non-deforming
regions. Bishop, Green and Hill 13 have shown that the possible deformable
region can be uniquely determined from one complete solution.
Discontinuities in stress and velocity in a complete solution are possible,
as previously discussed in sections 7.3.7 and 7.3.8, respectively. However,
it is emphasised that a stress discontinuity cannot occur across a slip line.
Also, Green 14 has illustrated that no velocity discontinuities can pass through
a singularity point.
A general rule relating to slip lines across which there is a tangential
velocity discontinuity, is that these slip lines either form a boundary of a
plastically deforming region or originate and/or terminate at a point of
stress singularity within a field. These conditions are generally necessary
to ensure that velocity compatibility is not violated.
Slip line fields have been proposed for plane strain deformation which
d~ not include a statically admissible extension of the slip line field into all
PLANE STRAIN PLASTIC DEFORMATION 209

{3

(a) Physical plane (b) Hodograph

Figure 7.23. A curvilinear element in a slip line field and verification that the
plastic energy dissipation in the element is positive

the non-deforming regions. Such solutions, if they are kinematically admis-


sible, are partial or incomplete solutions. They produce an upper bound for
the rate of energy dissipation and therefore overestimate the force required
to produce plastic flow. The lower and upper bound theorems will be
discussed in some detail in chapter 10.
The condition to be satisfied in order that the plastic energy dissipation
be positive has been given by Green 15 • However, the following method of
verifying that the condition is satisfied was suggested by Ford 16 •
In figure 7.23(a) is shown a curvilinear element ABCD bounded by a pair
of oc lines and a pair of p lines. The corresponding hodograph is given in
figure 7.23(b). The plastic energy dissipation in the element will be positive
provided the velocity of C relative to A represented by the vector, ~. in the
hodograph is positive, that is, having a sense from A to C, and the velocity
of D relative to B represented by the vector bd is positive, that is, having
a sense from D to B corresponding with the sign of the shear stresses in the
physical plane offigure 7.23(a).

7.4 SOME APPLICATIONS OF THE SLIP LINE FIELD


THEORY·

7.4.1 Plane strain inverted extrusion through an unlubricated square face


die with a fractional reduction R = 1/2
Forward or direct extrusion involves using a stationary container which is
closed at one end by a stationary die and at the other end by a moving ram or
plunger, as shown diagrammatically in figure 7.24(a). To minimise frictional
resistance, a close-fitting pressure pad is usually fitted at the face of the ram
thus enabling a clearance to exist between the ram and the container. The
workpiece material referred to as the billet is inserted into the container,
210 ENGINEERING PLASTICITY

Container Pressure pad

Extruded product
(a)

(b)

Figure 7.24. (a) Forward or direct extrusion; (b) backward or indirect extrusion

trapped between the die and the pressure pad and subjected to pressure by
the ram. The billet is caused to plastically deform mainly in the vicinity of
the die but first expands to completely fill the container before flowing out
through the die orifice to form the extruded product. As the extrusion
proceeds, the ram moves towards the die and the undeformed billet slides
forward in contact with the inner surface of the container. There is therefore
considerable frictional resistance at the interfaces between the billet and
container.
This frictional resistance, which is not easily defmed, can be avoided by
maintaining the billet stationary in a container closed at one end and causing
a close-fitting die to move against the free end of the billet as shown in figure
7.24(b). This process is known as backward or inverted extrusion. As the
extrusion proceeds the die moves forward into the container and the metal
is extruded backwards through the die and the hollow ram or plunger.
During inverted extrusion there is no relative movement between the
undeformed billet and the container and consequently no frictional resistance
at the interfaces.
However, it is convenient to assume that the die is stationary whilst the
container and billet move with unit velocity. During extrusion with a square
PLANE STRAIN PLASTIC DEFORMATION 211
Dead metal zone

Unit
H velocity

(a)

V=H/h=2

(b)

Figure 7.25. Plane strain indirect extrusion through an unlubricated square face
die (R= 1/2): (a) slip line field solution; (b) hodograph corresponding to top half of
extrusion

face die it is known that a dead metal zone forms in the corner between the
container and die as shown in figure 7.25(a). In this zone there is negligible
movement of material. The formation of a dead metal zone can be regarded
as a natural modification of the die profile converting it from a square face
die to a wedge-shaped die. Plastic flow occurs by intense shear strain at the
interface with the dead metal zone. If the billet is well lubricated so that
welding at the interface is prevented then the dead metal zone may be
separated from the end of the billet after extrusion. The dead metal zone is
usually curved in the regions where it meets the container and the die orifice
but approximately straight over the remainder of its length. As an approxi-
mation it is assumed that the face of the dead metal zone is inclined at an
angle of 45° to the geometric axis of symmetry. The exact angle of inclination
of this face is thought not to affect the analysis to any significant extent 17 .
Since the face of the dead metal zone meets the inner surface of the
container at an angle of 45° it constitutes a first slip line and satisfies the
212 ENGINEERING PLASTICITY

stress boundary condition for a frictionless container. The slip line field is
completed by two centre fan fields ABC and A'BC' emanating from the
points of stress singularity A and A', respectively, as shown in figure 7.25(a).
The exit slip lines AB and A'B intersect the axis of symmetry at angles of 45°
consistent with this plane being frictionless. Mathematically consistent
solutions for a single orifice, symmetric extrusion through a square face,
sharp-edged die were first presented by HiW 8 . All these solutions for frac-
tional reductions R < 0.88 assume the existence of a dead metal zone over
the die surface and a frictionless container. The slip line fields contain a
centre fan field similar to ABC with its centre located at the point of stress
singularity A at the corner of the die orifice.
The fractional reduction in area or, simply, fractional reduction which
characterises the degree of deformation attained during the extrusion is
designated by R and given by
R = change in cross-sectional area/initial cross-sectional area
= (A 0 - A)/A 0
= 1- (A/A 0 )
where A 0 is the initial cross-sectional area of the billet and A is the cross-
sectional area of the extruded product For the present case of plane strain
extrusion considering unit width of material : R = (H - h)/ H = 1 - (h /H) =
1/2. Also, for constancy of volume, Vh = H .1, where V is the velocity of the
extruded product leaving the die. Hence V = H jh = 2.
For the slip line field of figure 7.25(a) to be admissible, the velocity field
must be compatible with the velocity boundary conditions. These imposed
conditions are (a) the velocity across the entry slip lines BC and BC' must
everywhere be compatible with the velocity of the rigid billet; (b) no flow
occurs across the boundaries AC and A'C' of the dead metal zones; (c) the
velocity across the exit slip lines AB and A'B must be compatible with that
of the extruded product which is rigid on emerging from the die.
The procedure usually adopted to verify the compatibility of the chosen
slip line field with the velocity conditions is to construct the hodograph.
Because the slip line field is symmetrical about the geometrical axis of
symmetry for the container, die, etc., it is only necessary to consider one-half
of the field for purposes of analysis, say the top half. Therefore, only the
hodograph corresponding to this half of the slip line field need be constructed.
Referring to figure 7.25(a) any element of metal in the undeformed billet
to the right of the boundary slip line BC is assumed to move parallel to the
axis of symmetry with unit velocity. This is represented by a vector o!t of
unit length drawn from the origin 0 in the hodograph of figure 7.25(b). It is
assumed that there is a tangential velocity discontinuity along the boundary
slip line BC. This assumption is eventually shown to be valid by the hodo-
graph. An element crossing BC in the vicinity of C is therefore subject to a
velocity jump parallel to the tangent to BC at the point of crossing. For an
element crossing in the vicinity of C, this change in velocity is represented
PLANE STRAIN PLASTIC DEFORMATION 213
by the vector drawn from the point a, parallel to the tangent at C. After
crossing the line of velocity discontinuity, the element is constrained to move
parallel to the boundary AC of the dead metal zone so that its velocity
relative td the dead metal zone, which is stationary, is represented by a
vector drawn from the origin of the hodograph parallel to AC. This intersects
the vector ac in the point c, thus defining the magnitude of the velocity of
the element relative to the face of the dead metal zone, Ioc 1. and also the
magnitude of the tangential velocity discontinuity Iac I where Ioc I= Iac I=
(1/2)112.
Since the tangential velocity discontinuity along the entry slip line BC has
constant magnitude equal to Iac I, an element crossing BC at any other
point is subject to a velocity jump parallel to the tangent at the point of
crossing and has the same magnitude. Thus at any other point, such as D,
the tangent to the slip line BC at D is vertical and Iad I= Iac I· In the vicinity
of B, the tangent is at an angle of 45° to the axis of symmetry. The tangential
velocity discontinuity is ab and, consequently, the velocity relative to earth
at this point in the plastically deforming region is given by the vector Ob.
The element traverses the plastic region and emerges from it by crossing the
exit slip line AB where again it is subject to a tangential velocity discontinuity
in order to travel parallel to the axis of symmetry to form the rigid extruded
product.
The tangential velocity discontinuity across AB is represented by the
vector bd, and the direction of the velocity of the extruded product by the
vector oa produced. The intersection of these two vectors in the point, d,
defines the velocity of the extruded product ocl. It can be seen from the
hodograph of figure 7.25(b) that the velocity of the extruded product has a
magnitude of 2, thus showing that the chosen slip line field is compatible
with the velocity of an extrusion for a fractional reduction of 1/2.
It should be noted that the region covered by the slip line field is that in
which there is large plastic deformation and is in the vicinity of the die
orifice. The extruding billet is assumed to be long, so that when the slip line
field is established it remains stationary in space and is independent of time.
The extrusion is then referred to as being a steady state process.
The required extrusion pressure, p0 , can be determined from the slip line
field. It is convenient to start at the exit slip line AB since the stress
on the extruded material in the axial direction is known to be zero. Hence
at any point on the exit slip line AB, ax= 0. Also ryx = rxy = 0. From the
yield criterion of equation (7.12)

{(ax- ay) 2 /4} + r;Y = k 2


Therefore

This stress acts on a plane on which the shear stress ryx = 0 and is a principal
stress. Because the extrusion process is predominantly compressive it is
intuitively assumed that aY = - 2k and is compressive.
214 ENGINEERING PLASTICITY

The principal stress state at any point, P, on the exit slip line AB is then
given by a 1 = ax = 0, a 2 = a z = p = - k, a 3 = a Y = - 2k as shown in figure
7.26(a) and also by the corresponding Mohr stress circle diagram in figure
7.26(b). The direction of the intermediate principal stress, a 2 , is normal to
figure 7.26(a) but is not shown.
The direction of the algebraic maximum principal stress, a 1 , is contained
in the first and third quadrants of a right-handed rx-/3 coordinate system.
The rx and f3lines are therefore defined as shown in figure 7.26(a). The radial
straight slip lines, such as AB, are consequently defined as rx lines and the
circular arcs, such as BC, as f3 lines.
The principal stress state at the point P is equivalent to a hydrostatic
pressure, p, superimposed on a pure shear stress, k, as illustrated in figure
7.26(c). The senses of the shear stresses of magnitude, k, which occur on the a
and f3 slip lines are thus defined.

+k

(b)

-k
Oj=2k p:k p:k

p
(c) 0".=0
I

Figure 7.26. (a) Principal stress state at any point, P, on the exit slip line AB and
directions of the a. and fJ lines; (b) Mohr stress circle diagram for the stress state at
point P; (c) equivalent stress system at the point P
PLANE STRAIN PLASTIC DEFORMATION 215
It should be noted that the hydrostatic pressure is constant along the
straight a line, AB, such that PB = pA = k.
Along the fJ line BC, which is a circular arc, the angle ¢ varies hence the
magnitude of the hydrostatic pressure varies.
Using the second of the Hencky stress equations (7.33) B ---+ C along a
fJ line, p- 2k¢ =a constant. Therefore Pc- 2k¢c = PB- 2k¢B.
In moving along the entry slip line, BC, from B to C the tangent to the slip
line rotates through an angle ¢ = rc/2 anticlockwise and is therefore positive.
Hence Pc- 2k( + rc/2) = k, since </JB = 0 or Pc = k + krc = k{l + rc).
Along the face AC of the dead metal zone ACE, which is a straight slip
line, the hydrostatic pressure remains constant. The stresses acting on the
dead metal zone ACE are then as shown in figure 7.27(a), where q is the
normal die pressure acting on the die face AE and n is the normal pressure
at the interface between the frictionless container and the dead metal zone.

(a) q---t

+v

(b) .~ ~ 6
I
+V

(c)

Figure 7.27. (a) Stresses acting on the dead metal zone ACE; (b) components of
velocity at the point Q on the entry slip line BC; (c) the reference line for measurement
of the angle, </>, is chosen to coincide with the tangent to the f3line BC at the point C
216 ENGINEERING PLASTICITY

Resolving forces exerted on the dead metal zone ACE in the axial direction
gives q = Pc + k = k(n + 2) and if Pe is the extrusion pressure q(H/2) = PeH.
Therefore Pe = q/2 = k( n + 2)/2 or
Pe/2k = (n + 2)/4 ~ 1.29 (7.61)
The slip line field of figure 7.25(a) consists of two centre fan fields. For this
simple stress field, the Geiringer equations (1.41) can be used to verify that
the slip line field satisfies the velocity boundary conditions as follows :
All the IX lines are straight so that dcp is zero along the IX lines and from the
first of the Geiringer equations (7.41) du- vdcp = 0. Hence du = 0 and the
velocity along any IX line is constant. This is a property of any straight slip line.
Continuity of flow across the entry slip line, BC, at any point such as Q
requires that u2 + v2 = 1, since the billet is assumed to move with unit
velocity. If a particular IX line, say AQ, is inclined to the axis of symmetry at
an angle(}, then for constancy of volume the velocity on the IX line AQ,
- u = 1 cos(} (7.62)
as illustrated in figure 7.27(b).
The velocity of any point on the curved {3 slip line, BC, can be determined
by starting from the point C and knowing that v = 0 at all points on the face
AC of the dead metal zone ACE. Assume that cp = 0 coincides with the
direction of the tangent to the {3 line BC at the point C. This is then in a
direction (} = 7n/4, as shown in figure 7.27(c), since the angle(} is measured
anticlockwise from the axis of symmetry. Therefore (7n/4)- cp =e. Hence
f)+ cp = 7n/4 (7.63)
The second of the Geiringer equations (7 .41) for a {3 line is dv + u d cp = 0.
Therefore dv = - u dcp
dv
or dcp = - u =cos(}= cos { (7n/4)- cp}
Hence u= -cos{(7n/4)-cp} (7.64)
and dv =- u dcp =cos {(7n/4)- cp }dcp
Integrating v = sin {( 7n/4) - cp} +A
where A is a constant of integration.
At the point C on the dead metal zone, cp = 0 and v = 0. Therefore A =
- sin(7n/4) and
v =sin {(7n/4)- cp}- sin (7n/4) (7.65)
The velocity across the exit slip line, AB, must be compatible with the rigid
body movement of the extruded product. The tangent to the {3 line BC at
the point C rotates through an angle cp = - n/2(clockwise) to the tangent at
B. It follows that the velocity, v8 , across the exit slip line, AB, at B is therefore
given by
PLANE STRAIN PLASTIC DEFORMATION 217
vB = sin {( 7n/4) + n/2} - sin ( 7n/4)
= sin(9n/4)-sin(7n/4)
= (1/2)1/2 + (1/2)1/2 = 21/2 (7.66)
The same result is obtained for any other point along the exit slip line AB.
There is therefore a downward velocity of magnitude 21' 2 across the whole
of the exit slip line AB. This is a necessary requirement since the extruded
product to the left of this exit slip line is rigid.
There is a corresponding upward flow from the plastically deforming
region A'BC' with a velocity of magnitude 2 112 across the exit slip line A'B.
Thus, the net effiux velocity of metal across the whole exit boundary will
be the vector sum of these velocities. The resultant exit velocity in the axial
direction is
2vB cos(n/4) = 2.2 1' 2.(1/2) 1' 2 = 2 (7.67)
and the resultant velocity in the lateral direction is zero.
This is compatible with the motion of the rigid extruded product since,
for a fractional reduction R = 1/2, it has a velocity twice that of the billet
which was assumed to be unity.
The same result is obtained for any point on the straight exit slip line AB.
It can, however, be seen that the velocity component, uB, parallel to AB is
uB = - cos (} = - cos ( 3nj4) = (1/2 )1' 2. This is not compatible with the
velocity component of the rigid extruded product parallel to AB which is
2 cos(n/4) = 21' 2. However, this shows that a tangential velocity disconti-
nuity of magnitude (1/2) 1' 2 occurs across the exit slip line AB as represented
by the vector bd in the hodograph of figure 7.25(b).
The velocities calculated from the slip line field by using the Geiringer
equations have been shown to satisfy the velocity boundary conditions. The
chosen slip line field for an extrusion where R = 1/2 is therefore valid in this
respect. It will now be appreciated that it is usually simpler to construct the
corresponding hodograph to ensure that the proposed slip line field is
compatible with the velocity boundary conditions. It also remains for the
proposed slip line field to be tested to ensure that the energy dissipation is
everywhere positive in the plastically deforming regions, for example, by
employing the method suggested in section 7.3.11.

7.4.2 Indentation of a semi-infinite medium by a flat rigid punch


The indentation of a rigid-perfectly plastic, semi-infinite medium by a flat
rigid punch was first investigated by Prandtl 19 in 1920. The stress field
defined by the slip lines AKJHG in figure 7.28(a) combines regions of uniform
stress with centre fan fields.
This indentation problem is statically determinate since a unique stress
field AKJHGA can be obtained from the stress boundary conditions, namely,
the stress free surfaces AC and EG. A statically admissible extension of this
218 ENGINEERING PLASTICITY

St rus free
(a)

(b)

(c)

Figure 7.28. (a) Slip line field for the plane strain indentation of a rigid-perfectly
plastic, semi-infinite medium by a flat rigid punch; (b) hodograph assuming a dead
metal zone, CEJ, to be formed at the punch face (Prandtl solution); (c) alternative
hodograph for a frictionless punch (Hill solution)

stress field into the non-deforming region below AKJHG has been obtained
by Bishop 20 and Hill has shown that a block of fmite depth supported on a
frictionless plane may be considered semi-infinite if the depth is at least
8.75a 21 where 2a is the width of the flat punch.
It is assumed that there is a constant pressure over the face of the punch.
Only the case of incipient yield is considered because, as the plastic flow
PLANE STRAIN PLASTIC DEFORMATION 219
progresses, the shape of the plastically deforming boundary ABCDEFG
changes and it is then necessary to satisfy the boundary conditions on this
deformed boundary.
There is an infinity of admissible velocity fields that satisfy the boundary
condition for a normal velocity component on CE. From a theorem by
Bishop 20 it may be deduced that the plastically deforming region of any
velocity field cannot extend beyond the region AKJHGA. The velocity field
assumed by Prandtl involves a dead metal zone CJE and the hodograph for
this solution is given in figure 7.28(b). Only the right-hand half of the stress
field need be considered because of the symmetry of the field. If it is assumed
that there is no relative movement between the punch and the dead metal
zone CJE then no shear force is required to maintain the equilibrium of
the dead metal zone. The hodograph of figure 7.28(b) is then valid for any
punch roughness.
At the stress free surface EG, aY = -ryx = 0. From the yield criterion
{(ax- ay) 2/4} + -r;Y = k2
it follows that ax= ± 2k, since aY = 't"yx = 't"xy = 0.

Cal
+k

(b)

Figure 7.29. (a) Principal stress state at the point Gin the free surface, EG, and
directions of the IX and fJ lines; (b) Mohr stress circle diagram for the stress state at the
point G
220 ENGINEERING PLASTICITY

The punch pressure is compressive. Intuitively, it may be expected that ax


is compressive. It is therefore tentatively assumed that ax= - 2k on EG.
Since the shear stress, <yx' on the stress free surface, EG, is zero, this surface
is a principal plane and planes which are orthogonal to this stress free surface
are also principal planes. It follows that at the point G, a 1 = aY = 0 and
a 3 =ax= - 2k, as illustrated in figure 7.29(a) and hence a 2 = p = - k as
can be seen from the appropriate Mohr stress circle diagram given in figure
7.29(b).
It has been shown in section 7.3.9, that the slip lines intersect a stress free
surface at angles of± n/4. The direction of the algebraic maximum principal
stress, a 1 , is contained in the first and third quadrants of a right-handed rx-{3
system of coordinates. The rx and {3 lines are therefore designated as in
figure 7.29(a). It follows that GHJC is an rx line and EH, EJ, etc., are {Jlines.
In the physical diagram of figure 7.28(a) the rx lines are shown in outline and
the {3 lines are shown as dash lines.
The principal stress state at the point G is equivalent to a hydrostatic
pressure, p, superimposed on a pure shear stress, k, as illustrated in figure
7.30(a). The senses of the shear stresses of magnitude, k, which occur on the
rx and {3 slip lines are thus defined.

0"=0
I

G
o:•2k
3
• =

p=k p=k

(a)

+
p u n c h
2a

c E

(b)

Figure 7.30. (a) Equivalent stress system at the point G; (b) Stresses acting on
the dead metal zone CEJ
PLANE STRAIN PLASTIC DEFORMATION 221
The hydrostatic pressure at the point G on the a line GHJC is thus p0 = k
and since GH is a straight line there is no change in the angle c/J. Hence,
IJa=Pa= k.
However, HJ is a circular arc. There is therefore a change in the angle cjJ and
the hydrostatic pressure varies in magnitude from H to J. Applying the first
of the Hencky stress equations (7.33) in moving from H to J along the a line,
PJ + 2kcjJJ = Pa + 2k~ =a constant. Hence PJ + 2k(- n/2) = k since~= 0,
and PJ = k + kn = k(l + n).
Since the p line JE is a straight line there is no change in hydrostatic
pressure along the face JE of the dead metal zone. The stresses acting on the
dead metal zone CJE are then as shown in figure 7.30(b). The frictional stress,
-r, at the interface CE between the punch and the dead metal zone, which is
not shown, can vary from zero to a maximum possible value of k.
Let F be the punch force required per unit length in the z direction. By
resolving forces exerted on the dead metal zone in the direction normal to
the interface CE it can be shown that the indentation pressure, Pi, is
Pi= 2k{l + (n/2)} ~ 2k.2.57 (7.68)
and F = 4ka { 1 + (n/2)} (7.69)
A different solution proposed by Hill 22 assumes the punch to be frictionless.
The corresponding hodograph is then as shown in figure 7.28(c). However,
with this velocity field the deformation is confmed to the region DNPFD
when considering only the right-hand half of the stress field given in figure
7.28(a). In this case, the element ONE slides as a rigid body relative to the
frictionless punch face CE and with a velocity of magnitude 2 112 along the
surface DN. The velocity on FP is continuous and in the centre fan field
ENP, v = 0 and u = 2 112 • The triangular region EPF moves in the direction
of the surface PF with a velocity of magnitude 2 112 • In contrast with Prandtl's
solution the velocity field in the plastic zones is continuous. However, there
is an infinity of admissible velocity fields corresponding to the hodograph
of figure 7.28(c). These fields involve triangular elements of different sizes
sliding over the punch face CE.
This problem illustrates the non-uniqueness of solutions which refer to
a rigid-perfectly plastic material. For this reason, the construction of possible
stress fields and hodographs requires additional considerations and the
utilisation of experimental results.
It will be appreciated that the same hydrostatic pressure acts on the surfaces
EN and DN as acts on the surfaces EJ and CJ of the dead metal zone CJE
in the Prandtl solution. Consequently, the same uniform pressure acts
along the frictionless interface DE and hence the punch force F is given by
equation (7.69) as for the Prandtl solution.

7.4.3 Indentation of a semi-infinite medium with a wedge-shaped indenter


A slip line field solution for the indentation of a semi-infinite medium of a
222 ENGINEERI NG PLASTICITY

rigid-perfectly plastic material with an acute angled, wedge-shaped indenter


was proposed by Hill, Lee and Tupper 23 .
The smooth symmetrical wedge indenter of semi-angle, a, is forced
normally into the plane surface of the workpiece material. In the vicinity of
the indenter tip, three zones may be identified. Large plastic strains occur in
Zone 1. Zone 3 is that in which the material is elastically stressed and there-
fore treated hypothetically as a rigid zone whilst in the intermediate Zone 2
the material may be plastic but is restrained from moving. Only Zone 1 is
considered here and the slip line field corresponding to Zone 1 is shown in
figure 7.31(a) and the hodograph is presented in figure 7.31(b). The solution
allows for the 'piling-up' of material on each side of the indenter in a manner
such that geometrical similarity is preserved, that is, the stress field does not
change in shape but only changes in size as the penetration of the indenter
proceeds. The coronet or piled-up material EF A is equal in volume to that

Unit velocity

l
r-----~~-------~
Zone I (Slip line field)

E
I I
(a) / I
/ I
/
/ I
/ /

/
.," /
/

de

tfc'

(b)

Figure 7.31. (a) Slip line field for the indentation of a rigid-perfectly plastic,
semi-infinite medium by a smooth wedge-shaped indenter; (b) the corresponding
hodograph
PLANE STRAIN PLASTIC DEFORMATION 223
displaced by that part of the indenter GBF. However, a coronet is only
clearly observed during the indentation of a strain-hardening material.
Referring to figure 7.3l(a) it will be seen that the slip line field includes
the triangular region ABC which is a uniform compressive stress state. Since
the indenter face AB is frictionless, the slip lines AC and BC must meet the
surface AB at angles of n/4. ACD is a centre fan field with centre at A and
included angle (}which can be determined in terms of the indenter semi-angle,
a, from the condition of volume constancy. The triangular region ADE is
also a uniform compressive stress field.
The slip line BCDE is an a line and AC and AD are {3 lines which can be
defined following the method presented in the previous two sections. The
hydrostatic pressure distribution in the stress field ABCDEA can be deter-
mined by applying the Hencky stress equations (7.33) starting with the
knowledge that the surface AE is stress free. The hydrostatic pressure acting
on the surfaces AC and BC of the triangular region ABC is then shown to be
p = k(l + 2(}) (7.70)
The normal pressure, q, on the face AB of the wedge-shaped indenter is
assumed to be uniformly distributed and can be determined by considering
the equilibirum of the region ABC when it will be found that
q = 2k(l + (}) (7.71)
and the indentation force, F, is
F = 2q AB sin a = 4k( 1 + fJ) AH (7.72)
or the indentation force required per unit width of the indenter is
F = 4k(l + fJ) (7.73)
As the indenter semi-angle, a, approaches n/2 the angle (} also approaches
n/2 and the slip line field approaches that for indenting with a frictionless
flat punch given in figure 7.28(a). Equation (7.73) then reduces to equation
(7.69).
The solution for when the faces of the wedge-shaped indenter are rough
differs only in that the slip lines AC and BC do not intersect the face AB at
angles of± n/4 so that the slip line field is modified as shown in figure 7.32(a).
Results for various values of the coefficient of friction, J.l, at the interface
AB and of indenter semi-angle, a, have been given by Grunzweig, Longman
and Petch 24•
When the coefficient of friction, J.l, attains a critical value the frictional
stress at the interface AB attains its maximum possible value, k, that is, the
yield shear stress of the indented material. The appropriate slip line field is
that shown in figure 7.32(b).
The solution would be valid provided the slip line at the indenter tip
B intersects the face, AB, of the indenter at an angle of not less than n/4,
that is, LABC > n/4. If this condition is violated then the rigid material
224 ENGINEERING PLASTICITY

(a)
E

(b)
E

(c)

Dead metal zone

Figure 7.32. (a) Modified slip line field when Coulomb friction exists at the
interface between indenter and deforming material; (b) slip line field for indentation
with a wedge-shaped indenter when shearing of the workpiece material occurs at the
interface between the indenter and material; (c) slip line field for indentation with an
obtuse-angled wedge-shaped indenter when a dead metal zone is formed at the indenter
tip

in the vicinity of the indenter tip B is overstressed. This condition leads to


the possibility of a dead metal zone being formed at the indenter tip especially
for indenters with large included angles 2a and with high coefficients of
friction as illustrated in figure 7.32(c). If a= n/2 then the Prandtl solution
of figure 7.28(a) is applicable.

7.4.4 Cutting
The mechanics of the cutting of strip metal with plier-like tools can be
explained by extending the slip line field solution of figure 7.3l(a) for inden-
tation with a frictionless wedge-shaped indenter 21 .
A strip of finite thickness is indented by a pair of identical wedge-shaped
PLANE STRAIN PLASTIC DEFORMATION 225

(b)

ob

Figure 7.33. (a) Slip line field for cutting with a pair of opposed wedge-shaped
indenters (only top half shown); (b) corresponding hodograph for the right-hand half
of the stress field

indenters oppositely situated. This is represented in figure 7.33(a) which


shows the slip line field for only the top half since geometrical symmetry
exists about the horizontal axis which is then considered as a frictionless
foundation.
When the indenter first penetrates the strip metal the mode of deformation
is identical with that for indentation. 'Piling-up' of the material occurs at
the sides of the indenter but geometrical similarity is maintained and the
slip line field increases in size as penetration proceeds. However, because
the strip is of finite thickness, at some stage the plastically deforming region
extends through the whole thickness of the strip. Experimental evidence
indicates that a different mode of deformation then occurs. The material
ceases to 'pile-up' at the sides of the indenter but each half of the strip on
either side of the indenters moves outwards with a velocity, V, sliding
226 ENGINEERING PLASTICITY

horizontally over the frictionless foundation. The hodograph corresponding


to the right-hand half of the slip line field given in figure 7.33(a) is shown in
figure 7.33(b).
The indenting force required can be determined by choosing point, J,
as the starting point and first obtaining the magnitude of the hydrostatic
pressure PJ. The resultant force exerted on the portion of the strip to the
right of the slip line is zero. By considering the equilibrium of the forces acting
on the face AB of the indenter it can be established that ACEG and hence
ADFHJ are f3lines. If the region GHJH' is considered then the senses of the
shear stresses can be established since the plastic energy dissipation must be
positive when the point H separates and the point G must approach the
point J.
At any line element on the slip line ADFHJ which is inclined at an angle
¢ say, to the horizontal axis of symmetry the magnitude of the hydrostatic
pressure, p8 , is given by the second of the Hencky stress equations (7.33)
and is
Ps = PJ- 2k(n/4) + 2kcjJ (7.74)
The horizontal component of the total force due to the hydrostatic pressure
distribution over the slip line ADFHJ is

J: J:
Ps dh = {PJ- 2k(n/4) + 2kcjJ} dh (7.75)

and the horizontal component of the shear force acting over the slip line
ADFHJis

(7.76)

Since the resultant force in the horizontal direction is zero

J: {PJ- 2k(n/4) + 2k¢} dh- J: kdx = 0

Hence

{ (pJj2k) - ( n/4)} h + J: ¢ dh = x/2 (7.77)

When the hydrostatic pressure, pJ, has been determined from equation
(7.77), the Hencky stress equations can be used to determine the hydrostatic
pressure acting on AD and hence the indenting force required.
The mode of deformation considered is valid provided that the stress
normal to the foundation at the point, J, is compressive. The slip line field
of figure 7.33(a) may, however, be extended to include an isosceles triangle
PLANE STRAIN PLASTIC DEFORMATION 227
at the point J. The stress normal to the frictionless foundation is then zero
and that parallel to it is a tensile stress of magnitude 2k. The hodograph for
this case shows that the material in this triangular region moves vertically
upwards causing contact with the foundation to cease. This suggests that a
condition of tensile instability will occur with transverse necking leading to
fracture.
The effect of frictional resistance at the indenter faces on the critical
thickness at which the strip is parted in two has been considered by Johnson
and Kudo 25 . The deformation model may also be used to explain the cutting
of circular section wire with knife-edge pliers. Wire cutting has been discussed
at length by Johnson 26 and Mahtab and Johnson 27 .

7.4.5 Piercing
If a flat-faced punch with suitable clearances at the sides is envisaged as
indenting at the bottom of a deep vertically sided groove, the stress boundary
conditions are the same as for the initial penetration. A possible slip line
field for this process is that shown in figure 7.34 which resembles that of
figure 7.28(a) except that the fan regions centred on each of the punch
corners are developed through an angle n. The punch pressure required
for deformation at the bottom of the deep groove is given by
Pi= 2k( 1 + n)::::; 4.14. 2k (7.78)
that is, about four times the uniaxial yield stress of the deforming material
and hence the punch force per unit length is
F = 4ka( 1 + n) (7.79)

Figure 7.34. Slip line field for indentation at the bottom of a deep vertically
sided groove by a flat face punch
228 ENGINEERING PLASTICITY

This is therefore greater than that required for the initial indentation
given by equation (7.69).
The solution given by equation (7.79) is appropriate to a rough punch
with a dead metal zone formed at the face of the punch. The alternative
slip line field appropriate to a frictionless punch is shown in figure 7.34 by
dash lines. However, the mode of deformation suggested is only possible
if the material displaced by the punch is able to move upwards and out
behind the punch face.
The simplest form of piercing process is that where the billet is constrained
by a container of width, say 2b, which is comparable with the width of the
punch 2a. At some instant after the initial punch penetration, the deforming
material flows upwards past the punch and the process can be regarded as
steady state.

Unit velocity

2at

(a)

ac
(b)

Figure 7.35. (a) Slip line field for piercing with a flat face punch when the container
surface is frictionless and the fractional reduction= 1/2; (b) the corresponding hodo-
graph for right-hand half of the slip line field
PLANE STRAIN PLASTIC DEFORMATION 229
The slip line field solution presented in figure 7.35(a) is for a fractional
reduction of 1/2, that is, punch width/width of container = 2a/2b = ajb = 1/2
and assumes the container to be frictionless. In this case, the slip lines
intersect the container at angles of ± n/4 and the points A and A' are singular-
ities of stress. A dead metal zone ACA' is attached to the punch face and
moves downwards with the same speed as the punch. Considering the
right-hand half of the slip line field of figure 7.35(a), shearing of the
material occurs across the face AC of the dead metal zone and tangential
velocity discontinuities occur across the slip lines AB and CB. The corres-
ponding hodograph is shown in figure 7.35(b).
Material crossing the exit slip line AB moves as a rigid body and hence
the resultant force exerted on it is zero. The hydrostatic pressure exerted
on AB is thus equal to k. It can then be shown by examining the state of
stress at the point B that AB is a p line and CB is an a line. Using the first
of Hencky stress equations (7.33), the hydrostatic pressure at tha point C
and hence along the face AC of the dead metal zone is
Pc=k(1+n) (7.80)
The punch pressure required is
p = 2k{l + (n/2)} (7.81)
and hence the punch force required per unit length is
F = 4ka {1 + (n/2)} (7.82)
If there is frictional resistance at the container surfaces, the angles at which
the slip lines intersect the container would have to be adjusted accordingly.
As the ratio a/b decreases the punch pressure increases to the limiting value
for deep indentation of a semi-infinite medium given by equation (7.78).
It should now be appreciated that the mechanics of indentation, extrusion
and piercing are interrelated. The slip line field of figure 7.25(a) for inverted
extrusion with a fractional reduction of 1/2 is unchanged if the die is replaced
by undeformed billet material and the rigid extruded product is replaced by
a flat-faced punch moving in the opposite direction. There is consequently
a reversal of the geometric axis of symmetry and the container surface in
each process in relation to the geometry of the slip line fields of figures 7.25(a)
and 7.35(a).

7.4.6 Compression of a prismatic block between frictionless, rigid, parallel


platens for integral values of width/thickness ratio
Consider a rectangular section prismatic block of metal of thickness 2T to
be slowly compressed between smooth rigid, parallel platens of width 2W,
each moving with unit velocity a,s shown in figure 7.36(a). The overhang of
the metal on each side of the platens is assumed to be rigid.
Since the platens are frictionless the slip lines intersect the platens at
230 ENGINEERING PLASTICITY

Unit velocity

-vi
B
v--
---+ 2T

I. 2W
(a)
.I
J· v-
.,t:g'

(b)

Figure 7.36. (a) Slip line field for compression of a prismatic block between
smooth, rigid, parallel platens for integral values of W /T; (b) corresponding hodograph
for top right-hand quarter of the slip line field

angles of± n/4. Thus for integral values of W /T, the slip line field consists
of straight lines as shown in figure 7.36(a). When the platens move towards
each other with a relative velocity of 2, the material deforms as a series of
rigid triangular regions across which there is a tangential velocity disconti-
nuity of 2 112 and this is propagated by successive reflection from the platens
along their length. The hodograph corresponding to the top right-hand
quarter of figure 7.36(a) is given in figure 7.36(b). For all integral values of
WIT it can be shown that the normal platen pressure is uniformly distributed
and equal to 2k.
This solution is only valid for integral values of W /T because, for all
other values of W /T, the velocity discontinuity terminates on the exit
slip lines which is then incompatible with the rigid body motion of the
overhanging ends. However, a solution for non-integral values, 1 < W /T < 2,
has been obtained by Green 28 which is very complex and involves curved
starting slip lines. These solutions by Green have been analysed by Collins 29
employing a superposition property of slip line fields. An approximate
solution to this problem is given by Johnson and McShane 30 .
PLANE STRAIN PLASTIC DEFORMATION 231
7.4.7 Compression of a prismatic block between perfectly rough, rigid,
parallel platens
The compression of a rectangular section block of metal between perfectly
rough, rigid, parallel platens was apparently first investigated by Prandtl 31 .
However, it was shown by Hill, Lee and Tupper 32 that Prandtl's solution is
only valid if the workpiece material and the platens are of infinite width.
Nevertheless, if the block of metal is very wide compared with its thickness

2W
Unit 1 velocity

AI t le
- - 2T
-
I

cl 10
- +
Unit j vczlocity
(a)

p/2k
I

- - - - p/2k:2·871
I P. /2k::; 2·15
I y
1+(11"/21•1·29:
I

(b)
Figure 7.37. (a) Compression of a rectangular section block between perfectly
rough, rigid, parallel platens; (b) approximate slip line field and platen pressure distri-
bution for compression between perfectly rough, rigid, parallel platens when W/ T = 5.6
232 ENGINEERING PLASTICITY

the slip line field becomes more uniform with increasing distance from
either edge and approaches that of Prandtl's solution.
A prismatic block of metal which is compressed between perfectly rough,
rigid, parallel platens AB and CD is illustrated in figure 7.37(a). The width
of the platens is 2W and the thickness of the metal block is 2T such that
the width of the platens exceeds the thickness of the block, that is, W /T > 1.
Material of the block overhangs the platens at each end and this material is
assumed to be rigid. The platens approach each other with a relative velocity
of2.
One-half of the slip line field solution proposed by Hill et al., is given
approximately in figure 7.37(b) for various ratios of W /T::::;; 5.6. Since the
platens are perfectly rough, the slip lines are expected to meet the platens
normally and tangentially. Because symmetry exists only the left-hand half
of the slip line field is presented in figure 7.37(b).
The comers of the platens, such as A and C, constitute singularities of
stress and are convenient starting points in the construction of the slip line
field. From A and C, straight lines are drawn to intersect the horizontal
axis of symmetry in the point E at angles of ± n/4. The overhanging material
to the left of AEC moves outwards as a rigid body with velocity, say V,
and the resultant horizontal force exerted on this material is zero. The
hydrostatic pressure exerted on the exit slip lines AE and CE is therefore
equal to the yield shear stress k.
All the slip lines are required to meet the platens either normally or
tangentially. The field is therefore extended from A and C as radial lines and
circular arcs to produce the centre fan fields. The circular arcs meet the
platens normally in E 3 and E~ and the radial lines meet orthogonally on
the horizontal axis of symmetry and the platens tangentially. The centre
fan fields can then be extended as far as the point K corresponding to
W /T ~ 5.6. The simplest method is to use a small arc approximation which
approximates the curved element of a slip line by a chord that makes an
angle c/Jm for an IX line or c/Jm + (n/2) for a f3line with the positive x direction.
The angle c/Jm is the mean of the values of cp at the nodal points or at the ends
of the chord.
It should be noted that AEE~ is an IX line and CEE 3 is a f3 line. The centre
fan fields of figure 7.37(b) are subdivided into 15° arcs and the resulting
field is an equiangular mesh of 15° intervals. All slip lines must intersect
the horizontal axis of symmetry at angles of ± n/4. Angles measured anti-
clockwise from the x direction are considered positive and those measured
clockwise are therefore negative.
The extension slip line from the nodal point E 1 must rotate through an
angle of- 15° from E 1 to F. The nodal point F can be located by drawing
from the point E 1 a chord with a mean slope of c/Jm = - 37!0 , that is,
- (45° + 30°)/2 to intersect the horizontal axis in the point F. To extend
the field from a nodal point such as E 2 , a chord is drawn from E 2 with a
mean slope of c/Jm = - 22!0 , that is, - (30° + 15°)/2 since AE 2 is at - 15°
PLANE STRAIN PLASTIC DEFORMATION 233
from the x direction and at - 30° to the x direction at F 1 where it meets the
slip line of the f3 family. Then from Fa chord with a mean slope 4Ym = + 52! 0

that is, + (60° + 45°)/2 is drawn to meet in the point F 1 . The remainder of
the field can be constructed in a similar manner. However, because the mean
values of¢ are used, the slip lines will not meet orthogonally and the position
of the nodal points will not be exact. The slip line field can be drawn to a
greater degree of accuracy by using an equiangular mesh of less than 15°
intervals. The smaller the angular interval the more accurate the solution
will be although the amount of construction involved is considerably
increased. The true slip line field is constructed to a good degree of accuracy
by drawing pairs of orthogonal curved lines through the nodal points
defined by the intersection of the chords.
For convenience, the approximate slip line field of figure 7.37(b) has only
been extended as far as the point K for W /T R:; 5.6. Material above the slip
line G 3 K and below the slip line G~K, respectively, are dead metal zones
which are assumed to adhere to the platens and move with the same velocity.
A tangential velocity discontinuity thus occurs across these slip lines which
separate the dead metal zones from the plastically deforming region. Since
the precise stress distribution in the dead metal zones cannot be calculated
only the average stresses at the interfaces between the block and the platens
is determined.
At any point on the exit slip line AE, the zero principal stress in the hori-
zontal direction is the algebraic maximum principal stress and that in the
vertical direction is the algebraic minimum principal stress of magnitude
2k compressive. The intermediate principal stress is equal to the hydrostatic
pressure. The equivalent of this principal stress state is the same as that given
in figure 7.26(c). It follows that the stresses acting on the material above AE
are a hydrostatic pressure, p = k, and a shear stress k with a sense from A
to E. The family of slip lines radiating from A are thus e< lines and those
radiating from Care f3 lines as previously stated.
Moving along the f3line EE 3 from E to E 3 the rotation 4J = + n/4. Using
the second of the Hencky stress equations (7.33)
PE - 2k¢E = PE, - 2k¢E,
If it is assumed ¢E = 0 then ¢E, = + n/4 and since PE = k
PE, = k { 1 + (n/2)} 2.57k
R:; (7.83)
Moving from E 3 to F 2 along the e< line E 3 F 2 the rotation ¢ = - n/12.
Using the first of the Hencky stress equations (7.33)
PE, + 2k¢E, = PF2 + 2k¢F2
With ¢E, = 0 and ¢F2 = - n/12
PF 2 = k { 1 + (2n/3)} R:; 3.09k (7.84)
Moving from F 2 to F 3 along the f3 line F 2 F 3 the rotation ¢ = + n/12.
234 ENGINEERING PLASTICITY

·I
X
§3
I

T
I
------- K
(a)

V=s5•6---

(b)

Figure 7.38. (a) Stresses acting on a line element bs of the ex line G 3 K; (b) hodo-
graph corresponding to the approximate slip line field of figure 7.37 (b)

Using the second of the Hencky stress equations (7.33)


PF2 - 2kcjJF2 = PF, - 2kcfJF,
If cPF 2 = 0 then c/JF, = + n/12
Hence = k { 1 + (Sn/6)} ~ 3.62k
PF, (7.85)
In a similar manner it can be shown that Po, = k {1 + ( 7n/6)} ~ 4.66k
and the normal pressure at each point on the platen for the surface AG 3 is
determined. However, since the stress distribution in the dead metal zone
bounded by the slip line G 3 K is not calculable, only the average vertical
force or vertical stress at the interface with the platen can be determined.
Let {Js be an elemental length of the o: line G 3 K as shown in figure 7.38(a).
The vertical component of the force exerted on this line element is {JF =
(k sin cjJ + p cos cjJ) {Js where k is the shear stress, p the hydrostatic pressure
and cjJ is the angle of inclination of the line element to the x direction. The
total vertical component of the force exerted on the o: line G 3 K is then

F=k f sin cjJds+ f pcos cjJds = k f f


dy + pdx
PLANE STRAIN PLASTIC DEFORMATION 235
but p =Po, + 2k</J where <P is the magnitude of the rotation from G 3 to the
point considered on the slip line G 3 K.
Hence

F = kT + 2k I{(p 0 .J2k) + <P} dx

I
= k T + Po, dx + 2k I <P dx

or F/2k = (T/2) + (p 0 .J2k)X +I </Jdx (7.86)

where X is the horizontal distance of G 3 from the vertical axis of symmetry


and f<Pdx can be determined from figure 7.37(b). Note that this is the same
method described in section 7.4.4 for cutting.
If p is the mean normal pressure at the interface between the dead metal
zone and the platen, then

pj2k = T /2X + (p 0 .f2k) + I<P dx/X

I
= {(2 <P dx + T)/2X} + p0 .J2k (7.87)

For the approximate slip line G 3 K of figure 7.37(b) the following tabulation
can be performed:

Line

L\x 1.65 2.1 2.83


<P rad. 0.654 0.393 0.131
</J.L\x 1.08 0.825 0.371

Hence ~</J.L\x = 2.28, X= 6.58 and T = 2.5


From equation (7.87)
pj2k = ( {(2 X 2.28) + 2.5}/(2 X 6.58)] + 2.33
~2.87 (7.88)
The mean normal pressure on the platen, Pr• is given as the mean value
of (a) the constant pressure over the platen surface AE 3 , (PE)2k ~ 1.29);
(b) the steadily increasing pressure from E 3 to G 3 , (1.29 < pj2k < 2.33); and
(c) the mean normal pressure at the interface between the dead metal zone
and the platen, (p/2k ~ 2.87)
236 ENGINEERING PLASTICITY

Hence Py/2k ~ {(1.29 x 3.54) + 4.04(1.29 + 2.33)/2 + (2.87 x 6.58)} /14.2


~,., 15 (7.89)
It is interesting to note that, provided W /T;;:;, 1, the mean normal platen
pressure is given approximately by
Py/2k = (3/4) + (W /4 T) (7.90)
The normal pressure distribution at the platens when W /T ~ 5.6 is
shown in figure 7.37(b).
The hodograph corresponding to the approximate slip line field of figure
7.37(b) is given in figure 7.38(b). If the platens approach one another with
a relative velocity of 2 then at K there is a tangential velocity discontinuity
of magnitude 2112 at angles of± n/4. The velocity of the upper platen relative
to earth is represented by the vector ol
As material in the dead metal zone adhering to the upper platen crosses
G 3 K at K it is caused to move par~lel to the horizontal axis of symmetry
and its velocity is represented by ok. The remainder of the hodograph is
constructed in accordance with that for constructing tangential velocity
discontinuity fields which is discussed in chapter 10. However, the hodograph
to correspond with the true slip line field can be constructed so that it becomes
an orthogonal net of curved lines in accordance with the properties of the
slip line field.
The intermediate condition between frictionless or smooth platens and
perfectly rough platens is when Coulomb friction is present at the workpiece-
platen interfaces and the platens are partially rough. The effect of Coulomb
friction during the compression of a prismatic block has been considered
by Alexander 33 .

7.5 ESTIMATION OF PLANE STRAIN FORMING PARAMETERS


BY THE STRESS EVALUATION APPROACH
In this section, plane strain compression or forging, drawing, extrusion and
cold strip rolling of metal are considered with a view to deriving the approxi-
mate forming parameters by the stress evaluation approach. It is assumed
that a plane element of the workpiece material remains plane after deforma-
tion and other assumptions are made concerning the directions of the
principal stresses in the deforming material. Although these analyses are
approximate they provide useful relationships which are sufficiently accurate
for some purposes.

7.5.1 Quasi-static, plane strain compression between rigid, parallel platens


with Coulomb friction present at block-platen interfaces
Figure 7.39(a) indicates the stresses acting at any instant on a vertical plane
element of a rectangular section block of metal compressed between rigid,
PLANE STRAIN PLASTIC DEFORMATION 237
Normal platen prusure distribution
p !'Friction hill')

p:Yexp{piW-x)/T}

(b) Yup(pW /TI

I l
p
PI ate n y
-'Tyx=pp

I
O"x O",tdo,;
(a) 2T
X
dx

l PI a ten °1 p
IJP
I
X

}w
Figure 7.39. (a) The stresses acting on a vertical plane element during plane
strain compression between partially rough, rigid, parallel overhanging platens; (b) the
normal platen pressure distribution ('friction hill')

parallel overhanging platens which are partially rough. The frictional


stress ryx = J.lP is assumed to be low as may be the case during a cold forging
process. The coefficient of friction is J.l and p is the normal platen pressure.
The width of the block is 2 W, the thickness of the block is 2 T and its length
in the z direction is such that the deformation approximates to plane strain.
It is assumed that the stress ax is uniform across the section and is a principal
stress.
Considering unit length of the block in the z direction and resolving
forces exerted on the element in the x direction produces
(7.91)
for equilibrium.
This force equilibrium equation assumes that inertia forces are absent
and hence the compression is very slow, that is, quasi-static and that the
components of any body forces in the x direction are zero. Equation (7.91)
simplifies to
dax- (J.tpdxjT) = 0 (7.92)
Since ax is assumed to be a principal stress the normal platen pressure,
- p is also a principal stress even though a frictional stress exists at the
238 ENGINEERING PLASTICITY

interfaces between the block and the platens. However, if 'yx = f.lP is small,
the directions of the actual principal stresses will only be rotated through a
small angle from the directions of (Jx and - p.

(a) Normal platen pressure distribution and the 'friction hill'

From the yield condition which is independent of the choice of yield


criterion
(Jx- (- p) = Y (7.93)
where Y = 2k is the yield stress in plane strain. Therefore
d(Jx + dp = 0 (7.94)
or d(Jx = - dp (7.95)
Substituting for d(Jx in equation (7.92) and rearranging, dpjp = - f.ldx/T
which can be integrated to produce In p = - (J.lx/T) +A where A is a
constant of integration. However, at x = W, the vertical surface is stress free,
that is (Jx=O, and since (Jx+p= Y,p= Y. Hence In Y= -(J.1W/1)+A
Therefore ln(p/Y) = J.l(W- x)/T
or p/Y=exp{f.l(W-x)/T} (7.96)
The normal platen pressure thus increases inwards towards the vertical
geometric axis of symmetry exponentially from p = Y = 2k at the vertical
surfaces of the block where x = ± W to a maximum value at the axis of
symmetry where x = 0. Hence
Pmax/Y = exp(J.1W /T) (7.97)
The normal platen pressure distribution is shown in figure 7.39(b). If
the platens were smooth, that is, frictionless then the pressure distribution
would be uniform at p = Y = 2k. The effect of frictional resistance on the
normal platen pressure is therefore evident and for that reason the normal
pressure distribution is usually referred to as the ~friction hill'.
Since f.1 is low, the exponential term on the right-hand side of equation
(7.96) can be expanded as a series and truncated after the second term of
the series to give the approximation
pjY~l+{f.l(W-x)/T} (7.98)

(b) Force exerted by the platens and mean platen pressure


The force exerted by the platens on the vertical element per unit length in
the z direction is
dF = p dx ~ Y [ 1 + { J.l( W- x )/T} J dx
and the total force exerted by the platens on the block at any instant per unit
length in the z direction is
PLANE STRAIN PLASTIC DEFORMATION 239

F=2 f~pdxRi2Y J~[l+{JL(W-x)/T}]dx


Therefore FRi2WY{1 +(JLW/2T)} (~.99)

The mean platen pressure, p, is then given by


p=F/2WRi Y{1 +(JLW/2T)} (7.100)
During compression of the block the geometry of the plastically deforming
region changes continuously as the operation proceeds and the area of
contact between the block and the platens increases. The process is therefore
non-steady.
The progressive increase in the force required to maintain plastic defor-
mation of the block is not easily determined since this depends on the
spread of the block and the frictional conditions. However, it will be assumed
that the frictional conditions remain unchanged during compression.
Let 2 W0 and 2 T0 be the initial width and thickness of the block, respec-
tively. Assuming the block material to be incompressible then W0 T0 = W T.
The fractional reduction in thickness of the block is
(7.101)
Therefore T /T0 = 1 -Rand T = T0 (l - R) also W = W0 T0 /T = W0 /(l - R).
Substituting for Wand Tin equation (7.99) gives
F Ri { 2 W0 Y /( 1 - R)} [ 1 + {Jl W0 /2 T0 ( 1 - R )2 }] (7.102)
(c) Deformation energy
The energy required to compress the block at any instant per unit length
in the z direction is given by

E= f 2T 0
FdTRi2Y
f2To
{W+(JLW 2 j2T)}dT (7.103)
2T 2T

Substituting for W in terms of T into equation (7.103) and knowing that


T0 /T = 1/( 1 - R) produces
E Ri 2YW0 T0 [1n { 1/(1- R)} + (JLW0 /16T0 ){R(2- R)/(1- R) 2 }] (7.104)

(d) Plane of no slip


With reference to figure 7.39(a) it will be appreciated that the vertical plane
elements of the block which are initially situated on the right-hand side
of the vertical geometric axis of symmetry move in the positive x direction
during compression. However, those elements which are initially situated
on the left-hand side of the geometric axis of symmetry move in the negative
x direction. It follows that there must be an element which does not change
its position. This is situated at the geometric axis of symmetry which coincides
with the so-called plane of no slip. The plane of no slip will only coincide
240 ENGINEERING PLASTICITY

with the geometric axis of symmetry when the stresses acting on the extreme
vertical faces of the block are symmetrical, that is, of the same magnitude
and either both tensile or both compressive. Note that this condition is
satisfied for the case considered where the extreme vertical faces of the
block are assumed to be stress free.
It should be further noted that, for vertical plane elements initially on
the left-hand side of the plane of no slip, the frictional stress, ryx• at the
block-platen interfaces has opposite sense to that for the element shown
in figure 7.39(a) and is considered negative.

7.5.2 Quasi-static, plane strain compression between rigid, parallel platens


with constant frictional stress at block-platen interfaces
During compression of a block of metal at high temperature it is difficult
to maintain efficient lubrication at the block-platen interfaces. The frictional
stress, ryx• may therefore approach the maximum possible value which is
the yield shear stress, k, of the block material at the elevated temperature.
This limiting condition coincides with adhesion of the metal to the platens
when a thin boundary layer of the block adjacent to each platen surface
yields in shear. The regime is usually known as sticking friction or stiction.

(a) Normal platen pressure distribution


ryx, is assumed to exist at the block -platen
If a constant frictional stress,
interfaces, the equilibrium equation (7.92) can be written generally as
drrx- (mk dx/T) = 0 (7.105)
where ryx = mk and 0 < m < 1.
Again, assuming <Tx and - p are principal stresses and introducing the
yield criterion produces
dp= -mkdxjT (7.106)

which after integration leads to


p= Y + {mk(W- x)/T} (7.107)
Now, if the stiction regime is assumed to exist over the entire platen
surfaces, m = 1 and since 2k = Y
p = Y [ 1 + { (W- x )/2 T}] (7.108)
The normal platen pressure thus increases inwards in a linear manner
from p = 2k = Y at the vertical faces of the block, where x = ± W, to a
maximum value at the vertical axis of symmetry in figure 7.39(a) where
x = 0. The maximum value of the normal platen pressure, Pmax• for stiction
conditions is thus
Pmax = Y{1 + (W /2T)} (7.109)
PLANE STRAIN PLASTIC DEFORMATION 241
(b) Force exerted by the platens and mean platen pressure
The total force exerted by the platens on the block per unit length in the
z direction is

F=2 f~ pdx = 2 Y f~[ 1 + {( W- x )/2 T} J dx


Hence F = 2WY { 1 + (W/4T)} (7.110)
and the mean platen pressure, p, is given by
p=F/2W= Y{1 +(W/4T)} (7.111)

(c) Transition between Coulomb friction and stiction


However, the frictional conditions prevailing may not cause stiction over
the entire platen surfaces. In these circumstances, Coulomb friction can
then be present at the outer regions of the block where x approaches W.
As the normal platen pressure increases inwards, the magnitude of the
frictional stress <yx = .UP also increases. When this attains the value of the
yield shear stress, k, no further increase is possible. A region in proximity
to the vertical geometric axis of symmetry of the block will exist where the
stiction regime is applicable.
For these particular conditions, the equilibrium equation (7.92) can be
stated as dax- ( rdx/T) = 0 or
dp + (rdx/T) = 0 (7.112)
where r =.UP if .UP< k, r =kif .UP= k.
Let x = d be the distance of the point from the vertical geometric axis of
symmetry at which the transition occurs. This can be determined from the
condition .UP = k = Y /2 and substitution into equation (7.96). Hence
.UP= .uY exp {,u(W- d)/T} = k = Y /2
Therefore exp { .u( W- d)/T} = 1/2.u
or .u(W- d)/T = ln(1/2.u)
d= W -(T/.u)ln(1/2.u)
(7.113)
Substituting the value x = d from equation (7.113) into equation (7.96)
yields
[p/YJx=d = 1/2.u (7.114)
Integration of equation (7.112) gives p = - (kx/T) + B, where B is a constant
of integration, and this equation can be restated as pj2k = pj Y = - (x/2 T) +
C. When x=d,p/Y=1/2.u. Therefore C=(1/2.u)+(d/2T). Hence for
values of 0 < x < d
p/Y = (1/2.u) + {(d- x)/2T} (7.115)
242 ENGINEERING PLASTICITY

7.5.3 Quasi-static, plane strain drawing of a sheet material through a


wedge-shaped die
A wide sheet of metal which is drawn slowly through a wedge-shaped die of
small semi-angle, ()(, to reduce its thickness from 2H to 2h is shown in figure
7.40. It is assumed that the thickness of the sheet at any transverse section is
small compared with its width. The conditions then approximate to quasi-
static, plane strain deformation where displacements occur in the (xy) plane
and there is negligible displacement in the z direction. The axial stress is (Jx
at any vertical plane element in the plastically deforming region within the
die which is distant x from the exit of the die and where the thickness of the
sheet is 2y. The normal die pressure is p and the coefficient of friction is Jl..
With Coulomb friction present at the sheet -die interfaces, the frictional
stress is then r = Jl.P which is assumed to be low corresponding to good
lubrication being maintained. If back tension is employed, the axial stress
at entrance to the die due to this back tension is s. In certain respects, the
following analysis resembles that which is usually attributed to Sachs 34 .
Intuitively, it may be expected that (Jx is the algebraic maximum direct
stress and therefore the maximum principal stress (J 1 . Assuming that the
vertical stress, (JY' is uniform throughout a transverse section then it is the
algebraic minimum principal stress (J 3 .
Resolving forces exerted on the vertical plane element in the y direction
gives (JY dx + p ds cos()(- Jl.P ds sin()(= 0. Hence (JY dx + p dx- Jl.P dy = 0 or
(JY = - p + Jl.P tan()( (7.116)
In many cases, Jl.P tan()( is negligible compared with p and it may be assumed
that (JY = (J 3 ~- p.

Exit Entrance

Figure 7.40. The stresses acting on a vertical plane element during plane strain
drawing or extrusion through a wedge-shaped die
PLANE STRAIN PLASTIC DEFORMATION 243
(a) Drawing stress
Resolving forces exerted on the vertical plane element in the x direction
for unit width of sheet gives
(ux + duxHY + dy)- UxY + p ds sin a+ J.LP ds cos a= 0 (7.117)
for conditions of equilibrium. Noting that ds sin a = dy, ds cos a = dx =
dy cot a and neglecting the second order term dux dy, equation (7.117) after
simplification becomes
{ Ux + p(l + J.LCOt a)}(dyjy) +dux= 0 (7.118)
Under plane strain conditions, u 1 - u 3 = 2k = Y, which is independent
of the choice of yield criterion. Hence ux- (- p) = Y or
(7.119)
Substituting this value of pin equation (7.118), introducing the parameter
B = J.L cot a for convenience and rearranging
dux/{Bux- Y(1 +B)}= dyjy (7.120)
which can be rewritten as
(7.121)
where C =- Y(1 +B).
Equation (7.121) applies to wedge-shaped dies or to curved dies since
J.L, a and Y need not be independent of x. It can be applicable to a strain-
hardening metal since a relationship between Y and x could be deduced from
the true stress-natural strain (u,e) curve for the sheet metal. However,
assuming wedge-shaped dies and a rigid-perfectly plastic material, equation
(7.121) can be integrated with a, p. and Y constant and independent of x to
produce (1/B) In {Bux + C} =In y +In A, where In A is a constant of inte-
gration.
Therefore (7.122)
At entrance to the die where y = H, ux = s the stress due to back tension.
Hence (Bs + C) 1' 8 = HA or A= (Bs + C} 118/H. Therefore (Bux + C} 118 =
(y/H)(Bs + C} 118 and Bux + C = (y/H) 8 (Bs +C) or
ux = ( C/B) {(y/H)8 - 1} + s(y/H)8
= { Y(l + B)/B} {1- (y/H)8 } + s(y/H)8 (7.123)
At exit from the die, where y = h, ux = t which is the drawing stress. Therefore
t = { Y(l + B)/B} {1- (h/H)8 } + s(h/H)8 (7.124)
If back tension is not employed, s = 0 and
t= {Y(l +B)/B}{1-(h/H)8 } (7.125)
244 ENGINEERING PLASTICITY

Equation (7.123) shows that if back tension is employed, the drawing


stress is increased. However, this increase- is only a fraction of the stress due
to back tension.
(b) Maximum fractional reduction in a single pass
The limit to drawing occurs when the drawn sheet yields in uniaxial
tension assuming it to be free to contract laterally on leaving the die. The
uniaxial yield stress is shown by equation (6.57) to be equal to (3 1 12 /2) Y
where Y is the yield stress in plane strain.
The maximum fractional reduction in thickness of the sheet, Rmax, that
can be effected during a single pass can be determined from
t/Y = 3112 /2 = {(1 + B)/B} {1 - (h/H) 8 } + (s/Y)(h/H) 8 (7.126)
when back tension is employed or
t/Y = 3112 /2 = {(1 + B)/B} {1- (h/H)8 } (7.127)
if back tension is not employed.
The following typical values are assumed for the plane strain drawing
of a sheet metal through wedge-shaped dies without back tension:
,u = 0.05, IX = 10°, cot IX = 5.6713, B = ,u cot IX R; 0.2836, (1 + B)1B R; 4.53.
Thus t/Y = 3112 /2 = 4.53 { 1 - (h/H) 0 · 2836 }
1 - (h/H) 0 · 2836 = 3112/9.06 = 0.191
Therefore (h/H) 0 · 2836 = 1 - 0.191 = 0.809
(h/H) = 0.809 3 · 53 = 0.473
Hence Rmax = (2H- 2h)/2H = 1 - (h/H) = 1 - 0.473 R; 0.53
(7.128)
(c) Normal die pressure
The normal die pressure, p, can be determined by substituting for rJ x in
terms of p, that is, rJx = Y- p from equation (7.119) into equation (7.118)
and solving for pin a manner similar to that for determining (Jx. Alternatively,
since yield occurs when rJx- ( - p) = Y or p = Y- (Jx'
p= Y- [{Y(1 +B)}{1-(y/H)8 } +s(y/H) 8 J
which simplifies to
p = ( Y/B) { (y/H) 8 ( 1 +B)- 1}- s(y/H)8 (7.129)
The normal die pressure thus varies from p = Y - s at entrance to the die
where y =Hand (Jx = s, to a maximum value at exit from the die where y = h
and
p = ( Y /B) {( h/H) 8 ( 1 +B)- 1} - s( h/H)8 (7.130)
Equations (7.129) and (7.130) show that the normal die pressure is reduced
PLANE STRAIN PLASTIC DEFORMATION 245
in magnitude by employing back tension. It follows that the frictional stress,
r = pp, at the workpiece-die interfaces is reduced which leads to increased
die life. This is an important factor since dies are manufactured from expen-
sive high strength alloy steels and also the manufacturing costs are usually
high.

7 .5.4 Quasi-static, plane strain extrusion of a sheet material through a


wedge-shaped die
If the sheet of metal is extruded through a wedge-shaped die instead of
being drawn, the extrusion pressure required can be determined by integrat-
ing the equation of equilibrium (7.118) which leads to equation (7.122) that
is (Bax + C) 1i 8 = yA as for drawing, and substituting the appropriate
boundary stress values 35 .
At exit from the die, where y = h, ax = 0 since the extruded product is
stress free. Substituting these values in equation (7.122) produces
C 1 /B =hA
or A= C 118/h
Therefore (Bax + C)lfB = (y/h)Cl/B
and Bax + C = (y/h) 8 C
ax= ( C/B) {(yjh )8 - 1}
Hence ax=- Y {(1 + B)/B} {(y/h)8 -1} (7.131)
At entrance to the die, y =Hand ax= p., the extrusion pressure.
Therefore = Pe = - Y {(1 + B)/B} {(H/h)8 -1} (7.132)
and is compressive.
It should be noted that equation (7.132) for the extrusion pressure assumes
a frictionless container.

7.5.5 Redundant deformation during plane strain sheet drawing and


extrusion
The stress evaluation approach to estimate the drawing stress and extrusion
pressure is expected to produce underestimates. This arises because the
external constraint in the form of the die produces considerable internal
shear distortion of the workpiece in excess of that required to produce the
desired change of shape, that is, from thickness 2H at entrance to 2h at
exit from the die. Energy is therefore dissipated in producing this shear
deformation, which makes no useful contribution in effecting the desired
change of shape, and is referred to as redundant deformation.
Considering the case of drawing, an element of the workpiece near the
surface initially moves towards the die in a direction parallel to the horizontal
246 ENGINEERING PLASTICITY

~(-'----L.l...l:.
~- • EBB •
2H
HID • -+----If--

(a)

1--+---X

(b)

Figure 7.41. (a) Showing the shear distortion of an element as it passes through a
wedge-shaped die; (b) shear distortion of an element at a distance, y, from the horizontal
axis of symmetry

axis of symmetry but, on entering the die, is constrained to move parallel to


the die surface and hence has an inward velocity component in they direction.
This produces an internal shear deformation of the element. As the element
emerges from the die exit, it is again subjected to shear deformation so that
it proceeds once more in an axial direction within the drawn sheet. An
element initially on the axis is not subjected to redundant deformation.
The shear distortion of an element as it passes through a wedge-shaped
drawing die is shown diagrammatically in figure 7.41(a).
It follows that different elements in the plastically deforming region
within the die are subjected to different degrees of redundant deformation
varying from zero at the axis to a maximum shear distortion at the die
surface. The energy dissipated in producing redundant deformation is
therefore greater when the die semi-angle, oc, is large than when the die is
long with a small semi-angle. However, long dies involve greater frictional
resistance. Since the energy dissipated in overcoming frictional resistance
and producing redundant deformation are both superimposed on the energy
required for homogeneous deformation, there is an optimum die angle for
PLANE STRAIN PLASTIC DEFORMATION 247
sheet drawing giving the least drawing stress for a given fractional reduction
in thickness of the sheet. Extrusion dies usually have a large included angle,
for example, 2oc = 120°. The energy dissipated in producing redundant
deformation during extrusion is therefore greater than in drawing.
At entrance to the die, during plane strain sheet drawing the element at
distance y from the axis is dx long and dy thick as shown in figure 7.41(b).
It is sheared through an angle 8. The angle () varies from zero at the axis
to a maximum of oc at the die surface. Assuming this variation to be linear
() = (y/H)oc (7.133)
Energy dissipated per unit volume in shearing the element
= k() = (ky/H) oc (7.134)
where k is the yield shear stress of the material. The volume of the element
is dx.dy per unit length in the z direction. Therefore the energy required to
shear the element
dE= (ky/H) ocdxdy (7.135)
and the total energy required to produce the redundant deformation at
entrance to the die

E = (2kocdx/H) f>dy

= (2kocdx/H)(H 2 /2)
=kHocdx (7.136)
The volume of material entering the zone R; 2H dx. Therefore energy
dissipated per unit volume in producing redundant deformation at entrance
to the die
R;kH oc dxj2H dx
R; koc/2 R; Y oc/4 (7.137)
for a rigid-perfectly plastic material where Y is the plane strain yield stress.
Equation (7.137) becomes
Yoc/2.3 112 (7.138)
for a rigid-perfectly plastic material where Y is the uniaxial yield stress. For
a strain-hardening material equation (7.137) becomes
Yocj2Yi 2 (7.139)
where Yis the mean yield stress.
It should be noted that the plane strain yield stress (constrained yield
stress)= 2/3 112 times the uniaxial yield stress (unconstrained yield stress).
Similar expressions can be deduced for the energy dissipated per unit
248 ENGINEERING PLASTICITY

volume in producing the redundant deformation at exit from the die when
the elements are again sheared to move parallel to the axis within the drawn
Product.
The additional stress, ared• required to produce redundant deformation
during plane strain drawing or extrusion is thus given approximately by
O"red = Yr:t./2 (7.140)
for a rigid-perfectly plastic material where Y is the plane strain yield stress or
O"red = Y r:t./3112 (7.141)
if Y is the uniaxial yield stress, and
O"red = v-,
~v. 13
112 (7.142)
for a strain-hardening material if, Y, is the mean yield stress.

7.5.6 Quasi-static plane strain cold rolling of plate and strip metal
Rolling is the process of reducing the cross-section of the workpiece by
passing it between two rotating rolls. The reduction in cross-section is
accompanied by elongation in the direction of rolling and there may also
be lateral spread of the workpiece.
Large reductions in cross-section as, for example, in the rolling of ingots
and billets are achieved at elevated temperatures, that is, above the recrystal-
lisation temperature such that the homologous temperature T" > 0.5. This
process is therefore usually referred to as hot rolling. Hot rolling is one of
the major industrial methods of producing bars of rectangular cross-section
and the hot rolling mill consists of two large parallel cylindrical rolls mounted
vertically orie above the other. Vertical edging rolls may be provided to
control the width of the workpiece during rolling. The mill is consequently
described as a 2-high mill as distinct from a 4-high or more complex mills
which are used in cold-rolling. Rolls having special profiles are used in the
hot-rolling of other cross-sections including round, hexagon, channel, angle
and I section.
During rolling, the plastically deforming region is restricted to a zone of
small volume. It is thus possible to process large ingots using mills of
moderate capacity. For steel sheet manufacture, ingots may weigh more than
0.2 MN and have a cross-section of 600 mm square. The rolling operation
is fast and is more economical than forging. A limiting factor in speed of
manufacture is the time required to transport the slab back to the entry
side of the rolls. However, this limitation is obviated in most cases by using
reversing mills. This is important because in hot working processes the
workpiece hardens rapidly on cooling. Hot rolling improves the mechanical
properties of the cast metal by homogenising and refining the structure
producing greater strength and toughness.
Wide slabs combined with high roll force may cause the rolls to deflect
PLANE STRAIN PLASTIC DEFORMATION 249
due to bending so that the rolled product becomes thicker in the middle
than at the edges. This had led to the use of 4-high mills in which the work
rolls are of small diameter, to reduce the roll force, supported by larger
diameter back-up rolls.
As the cross-section ofthe workpiece approaches the shape ofthe required
product, the final rolling operation may be performed at room temperature.
It is then referred to as cold rolling. Cold rolling, as in other metal forming
processes carried out at room temperature, produces an improved surface
finish, good dimensional accuracy and also improved mechanical properties.
However, greatly increased roll force, roll torque, energy and power are
required for identical reductions and roll speeds compared with hot rolling.
Aluminium foil is rolled in 2-high mills even to thicknesses of the order
of0.02 mm, as for cigarette and chocolate wrapping, which is folded over and
rolled double. The rolls are screwed down so that the roll gap is completely
closed and, in the absence of the foil, may be under pressure initially which
is equivalent to a negative roll gap. The thickness of the foil is then controlled
by tension applied to the power driven coiler.
For the cold rolling ofthin strip and foil of hard materials, such as titanium
alloys or stainless steel for the manufacture of razor blades, a number of
mills of the cluster type have been introduced. These include the Sendzimir
mill 36 and the planetary mill, which was also developed by Sendzimir, in
which the back-up rolls are themselves supported by work rolls of small
diameter.
The published literature concerned with rolling is extensive. No attempt is
made here to refer to all the reported experimental and analytical work on the
subject. However, detailed discussions on rolling theory are to be found in
the textbooks by Underwood 3 7 , Larke 38 , Alexander and Brewer 39 , Avitzur40,
Rowe41 and Tarnovskii, Pozdeyev and Lyashkov 42 • An English translation
of the book by Tselikov and Smirnov43 is concerned with rolling mills.
A survey of the literature shows that analyses have been mainly related
to rolling of strip, plate and sheet metal and no exact solutions are available
at present. The original analysis employing a slab method of solution to
obtain the differential equations of equilibrium and determine the stress
distribution was apparently first introduced by von Karman 44 in 1925.
He also recognised that the frictional stresses at the interface between the
workpiece and the rolls change sense because of the change in relative
velocity between the rolls and the workpiece as it passes through the rolls.
Orowan45 analysed the rolling process based on the concept of channel
flow in order to assume a spherical stress state and formulated an approxi-
mate method of allowing for redundant deformation. Bland and Ford46
have shown that the allowance required for redundant deformation is very
small when the arc of contact between the rolls and the workpiece is greater
than the mean thickness of the strip. The reductions effected in industrial
practice during a single pass are such that the arc of contact is usually greater
than three times the strip thickness.
250 ENGINEERING PLASTICITY

As rolling proceeds, the rolls are themselves elastically deformed over the
arc of contact with the workpiece referred to as roll flattening and should be
taken into consideration when determining the roll force and roll torque.
Roll flattening was first discussed by Hitchcock 47 and the rolling of thin strip
with reference to roll flattening was studied by Stone48 . The maximum
reduction which can be achieved in cold strip rolling has been discussed by
Hill 49 , Ford 50 and Avitzur 51 .
Slip line field solutions for the rolling problem have been presented by
Alexander 52 , Alexander and Ford 53 , Druyanov 54 and Firbank and
Lancaster 55 . An upper bound solution for fast hot rolling has been given
by Johnson and Kudo 56 .
In hot-rolling, the flow stress of the workpiece material is strain-rate
dependent and the frictional stress at the interface between the rolls and the
workpiece is high. However, in cold rolling of plate and strip the flow stress
is nearly independent of the strain-rate and the frictional stress is relatively
low. The subject of this section is therefore restricted to the quasi-static,
cold rolling of plate or strip metal where the width to thickness ratio of the
workpiece is, say, not less than ten. Since the width of the strip is large
compared with the length of the arc of contact between the rolls and the
strip, the constraint of the non-plastically deforming material inhibits lateral
spread which normally does not exceed 2 per cent. The deformation is then
essentially plane strain.
The main assumptions made in the theory can be summarised as: (1) the
workpiece material is rigid-plastic strain hardening; (2) the arc of contact
is circular; (3) the deformation is plane strain; (4) the coefficient of friction
is constant over the arc of contact; (5) plane sections normal to the direction
of rolling remain plane.
Consider a metal strip of width, b, and initial thickness, H, to enter the
gap between two parallel cylindrical rolls as shown in figure 7.42(a). The strip
is first compressed elastically until it yields. It is then plastically deformed
but strain hardens with increasing strain and, on emerging from the roll gap,
elastic recovery occurs such that the final thickness of the strip is h.
It will be assumed that the strip material is rigid-plastic strain hardening.
The elastic deformation of the strip is therefore considered to be insignificant
and the effect on the magnitudes of the roll force and roll torque neglected.
This imposes a limitation on the validity of the theory since, for low fractional
reductions in thickness and when rolling very hard strip metals, the elastic
strains are not small compared with the plastic strains.
As rolling proceeds, the rolls themselves are distorted elastically over the
arc of contact. It will be assumed that the arc of contact is circular and of
radius R', which is greater than the radius R of the undeformed rolls.
In passing through the roll gap, the strip reduces in thickness from H to h
and the velocity of the strip increases steadily from V at entry to v at exit.
The linear velocity of a point on the cylindrical surface of a roll must have
some value between V and v. It follows that since the deformation is plane
PLANE STRAIN PLASTIC DEFORMATION 251

Radius of und~form~d
roll= R

h
--
v

.,,, , ... --1


_J Limits of th~
1 plastic zon~

(a)

Entry sid~
pR'dct>sinct>
ct>
pRdct>cosct>

Figure 7.42. Cold rolling of strip metal: (a) stresses acting on the plastically
deforming material in the roll gap within the arcs of contact; (b) forces exerted on
elements on the entry and exit sides ofthe neutral plane

strain and the material is incompressible


bhv = byu = bHV (7.143)
where y and u are the thickness and velocity ofthe strip at some intermediate
position in the roll gap.
On the entry side of the roll gap, the rolls move faster than the strip and
the frictional force draws the strip into the rolls. At the exit side of the roll
252 ENGINEERING PLASTICITY

gap, the strip moves faster than the rolls and the frictional force tends to
oppose the motion of the emerging strip. Consequently, at some intermediate
plane, referred to as the neutral plane, the strip and rolls have the same linear
velocity. The position of the neutral plane depends on the magnitudes of
the stresses due to back and front tensions, t 1 and t 2 , respectively, applied
to the strip.

(a) Normal roll pressure


The stresses acting on the element in the roll gap on the entry side of the
neutral plane are shown in figure 7.42(a). The normal roll pressure is p and
the frictional stress is p.p where p. is the constant coefficient of friction. For
unit width of strip, the forces exerted on elements on both the entry and exit
sides of the neutral plane are shown in the views of figure 7.42(b).
Resolving forces exerted on an element in the direction of rolling for unit
width of strip
(ax+ dax)(y + dy) -lTxY + 2p.pR'd¢cos¢ + 2pR' d¢sin¢ = 0 (7.144)
for the condition of equilibrium.
The negative sign applies to all elements between the entry plane and the
neutral plane and the positive sign applies to all elements between the neutral
plane and the exit plane. From equation (7.144), ydax + axdy + 2pR'd¢
(sin¢+ cos¢)= 0 if second order quantities are neglected. Rearranging
d( axy )/d¢ = - 2p R' (sin¢+ p. cos¢) (7.145)
In cold rolling, the arc of contact is small, usually less than 6°, therefore
sin ¢ ~ ¢ and cos ¢ ~ 1. Therefore
d(axy)/d</J =- 2pR'(</J + p.) (7.146)
Resolving forces exerted on an element in the direction normal to the direc-
tion of rolling for unit width of strip
- pR'd¢ cos¢+ p.pR' d¢ sin¢+ aydx = 0 (7.147)
for the condition of equilibrium. Therefore
-pR' d¢ cos ¢(1 ±tan¢)+ aydx = 0
or aY = p(1 ±tan¢) (7.148)
since R' d¢ cos ¢ ~ dx
However, since ¢ is small, tan ¢ ~ 1. The normal roll pressure, p, is thus
approximately equal to the vertical stress aY. Equation (7.148) can then be
written as
(7.149)
with an error of less than 1 per cent.
Assuming ax and p to be principal stresses then the yield condition which
is independent of the choice of yield criterion adopted for plane strain
PLANE STRAIN PLASTIC DEFORMATION 253
deformation gives <Jx- (- p) = 2k where k is the yield stress in pure shear.
According to the von Mises yield criterion 2k = (2/3 1 12 ) Y where Y is the
yield stress in uniaxi<d compression and for the Tresca yield criterion k = Y /2.
(Jx = 2k- p
or d<Jx/dcf> = d(2k- p)/dcf>
Therefore d( <JxY)/dcf> = d {y(2k- p)} /d 4>
=d[2ky{1-(p/2k)}]/d¢
Hence d( <JxY)/d¢ = d [ 2ky { 1 - (p/2k)} ]/de/>= - 2pR' ( 4> + J.l)
or - d(<JxY)/dcf> = d [2ky{ (p/2k)- 1} ]/de/>= 2pR'( 4> + J.l)
which can be written as
2ky { d(p/2k)/d¢} + {(p/2k)- 1 }{ d(2ky)/dcf>} = 2pR'( 4> +J.l) (7.150)
Since the strip strain-hardens as rolling proceeds, the yield stress 2k
increases as y decreases through the roll gap. The term { (p/2k)- 1}
{d(2ky)/d¢} is thus negligible compared with the term 2ky{ d(p/2k)/d¢}
and 2ky tends to remain constant. The coefficient {(p/2k)- 1} is also small
for cold-rolling. Equation (7.150) can be reduced to
2ky{ d(pj2k)/dcf>} ~ 2pR'(cf> + J.l) (7.151)
or d(p/2k)/dcf> ~ 2pR'( 4> + J.1)/2ky
and {d(p/2k)/dr/>} /(p/2k) ~ 2R'( cf> + J.l)/y (7.152)
The variation in strip thickness y through the roll gap is given by y =
h + 2R' ( 1 - cos 4>). However 1 - cos 4> ~ ¢ 2 /2 so that
y ~ h + R'c/> 2 (7.153)
Substituting this value for yin equation (7.152) produces
{ d(p/2k )/de/>} /(p/2k) ~ 2R' ( 4> + J.1 )/(h + 2R' ¢ 2)

~2(¢ +J.l)/{(h/R') + ¢ 2 }
~ 2¢/ {(h/R') + ¢ + 2J.1/ { (h/R') + ¢
2} 2}

Integration produces
ln(p/2k) ~In {(h/R') + ¢ 2 } +2J.l(R'/h) 1 tan- 1
1 2

{(R'/h) 112 cf>} +InC


where InC is a constant of integration.
Let Q = 2(R'/h) 112 tan- 1 {(R'/h) 112 ¢} then
In (p/2k) ~In (y/R')+ J.lQ +InC
or pj2k ~C(y/R')exp( +J.lQ) (7.154)
254 ENGINEERING PLASTICITY

The dimensionless roll pressure for the entry side of the neutral plane is
thus given by
p- /2k R:< c-(y/R')exp(- J.1Q) (7.155)
and for the exit side
p+ /2k R:< c+(y/R')exp( + J.1Q) (7.156)
Let k 1 and k2 be the values of the yield shear stress at entry and exit,
respectively. At exit, ax= t 2 , the stress due to front tension. Therefore
t 2 - ( - p+) = 2k 2
or p+ = 2k 2 - t2 (7.157)
Also at exit y =hand since ¢ = 0, Q = 0. Therefore p+ /2k 2 = c+ (h/R') and
substituting for p+ from equation (7.157)
p+ /2k 2 = (2k 2 - t 2 )/2k 2 = 1- (t 2 /2k 2 ) = c+(h/R')
Hence c+ = (R'/h){ 1- (t 2 /2k 2 )} (7.158)
and the roll pressure on the exit side is given by
p+ R:< (2kyjh){ 1- (t 2 /2k 2 )} exp( + J.1Q) (7.159)
On entry side, ax = t 1 , the stress due to back tension. Therefore t 1 - ( - p-) =
2k1
or p- = 2k 1 - t1 (7.160)
andy= H. Therefore (2k 1 - t 1 )/2k 1 = c- (H/R')exp(- J.1Q 1 )
where Q1 is the value of Qat entry.
Hence (7.161)
The dimensionless roll pressure for the entry side of the neutral plane is
thus given by
p- /2k R:< (R'/H){ 1- (ttf2k 1 )} exp( + J.1Q 1 )(y/R')exp(- J.1Q)
The roll pressure on the entry side is then given by
(7.162)

(b) Position of the neutral plane


The position of the neutral plane can be determined since, at the neutral
plane, p+ = p-. That is, combining equations (7.159) and (7.162)
(2kyn/h ){ 1 - (t 2 /2k 2 )} exp ( + J.1Qn) = (2kyn/H){ 1 - (ttf2k 1 )} exp (Q 1 - Qn)
(7.163)
where Yn is the thickness of the strip at the neutral plane and Qn is the value
of Q when ¢ = ¢nat the neutral plane.
PLANE STRAIN PLASTIC DEFORMATION 255
Simplifying equation (7.163) produces
exp {Jl(Q 1 - 2Qn)} = (H/h){1- (t 2 /2k 2 )}/{1- (t 1 /2k 1 )}
Therefore
J1(Q 1 - 2Qn) = ln [ (H/h ){ 1 - (t 2 /2k 2 )} /{ 1 - (td2k 1 )} J
and Qn = (Qd2)- (1/2/1) ln [(H/h){1 - (t 2 /2k 2 )} /{1 - (td2k 1 )}] (7.164)
Since Qn = 2(R'/h) 1 12 tan- 1 {(R'/h) 112 <f>n}
tan - 1 { ( R'/h ) 1 12 <f>n} = ( h/ R') 112 Qn/2
Hence <Pn = ( h/R') 112 tan {( h/R') 112 Qn/2} (7.165)
The value of <Pn can thus be determined by substituting for Qn from
equation (7.164). If front and back tensions are not employed then equation
(7.164) reduces to
Qn = (Qd2)- (1/211) ln(H/h) (7.166)
The normal roll pressure increases from both the entry and exit sides
to attain a maximum value at the neutral plane. The normal roll distribution
curve over the arc of contact is therefore hill shaped and, as before, is referred
to as the 'friction hill'. Normal roll pressure curves which can be derived from
this theory for the cold-rolling of strip metal have the form shown in figure

Some pass without


employing front
and bock tensions

t
"'...:::>
.....
Ul
Ul

Q.

...0 With bock


0
tension t 1
...e0
z

<Pn Angle cf>


-
Figure 7.43. Normal roll pressure curves showing the effect of employing front
and back tensions to the strip metal
256 ENGINEERING PLASTICITY

7.43. Two cases are illustrated, namely, when front and back tensions are
employed indicated by the lower curves; the upper curves are for the case
when front and back tensions are not employed for the same fractional
reduction in thickness of the strip.

(c) Rollforce
The yield stress, 2k, varies as the strip passes through the roll gap and
therefore varies with the angle c/J. Values of the yield stress for the strip
material corresponding to different fractional reductions can be determined
experimentally using the plane strain compression test devised by Ford 57
and later developed by Watts and Ford 5 8 . The plane strain compression test
has been described in some detail in section 6.3.2.
With the variation in yield stress and also the coefficient of friction known,
the roll force per unit width of strip can be determined by integration of the
normal roll pressure over the arc of contact.
Thus, the roll force per unit width of strip is given by

F= f: 1
pR'dcp

or F ~ R' { J:" p+ d cp + J:>- d cp} (7.167)

It has been found from some experiments that the yield stress curve as
determined from the plane strain compression test produces an under-
estimate of the roll force. This is probably due to redundant deformation
in cold-rolling causing the strip metal to strain-harden more rapidly than
in plane strain compression. Furthermore, the yield stress for the high speed
rolling of strip metal could be expected to be higher than that predicted by a
quasi-static plane strain compression test. Comparisons between theory
and experimental result are given in the papers by Hessenberg and Sims 59
and Whitton and Ford 60 •
The roll force is decreased if the frictional stress at the interface between
the rolls and the strip is reduced to the minimum or if front and back tensions
are employed. The effect of using five different lubricants on the magnitude
of the roll force when rolling thin titanium strip was investigated by Wilcox
and Whitton 61 . It is suggested that the roll force can be reduced by as much
as 60 per cent. Bedi and Hillier 62 have analysed the hydrodynamic effect
of oil between the rolls and the strip metal on roll force and roll torque.

(d) Roll torque


The mean torque for each roll per unit width of strip is given by the integral
of the moment about the roll axis of the friction force along the arc of contact.
Unless there is considerable roll flattening, the moment of the components
PLANE STRAIN PLASTIC DEFORMATION 257
of the normal forces may be neglected. The moment arm of the frictional
forces is the distance from the roll axis to the roll surface which is approxi-
mately the radius, R, of the undeformed roll. Since the frictional stress has
opposite sense on the entry and exit sides of the neutral plane, the resultant
roll torque, T, per unit width of strip is given by

T ~ R{J:~ JlP- R' dq?- J:~p+ R' dq?}

or (7.168)

Equation (7.168) involves the difference of two large quantities of nearly


the same magnitude. It is therefore difficult to determine the roll torque
with sufficient accuracy by using equation (7.168). However, it was shown
by Bland and Ford46 that an alternative equation can be derived by consider-
ing the equilibrium of the forces exerted on the strip in the roll gap in the
direction of rolling.
Substituting the yield condition, ux = 2k- p into equation (7.146) gives
d(uxy)jdq? = d{(2k- p)y }fdq? = - 2pR'(q? +Jl)
Integrating between the limits of q? = 0 at exit and q? = q? 1 at entrance to
the roll gap

t 1 H- t 2 h = 2R'{f>q?dq?- Jl(J:>- dq?- f>+ dq?)}

Therefore Jl(J:>- dq?- J:• p+ dq?) = J:~q? dq? + { (t 1 H- t 2 h )/2R'}

If this is now substituted into equation (7.168) the equation for the roll
torque becomes

(7.169)

(e) Determination of the coefficient offriction


The following method for determining the coefficient of friction, Jl, during
the cold rolling of strip metal, by direct measurement, was originally
suggested by Bland and studied by Whitton and Ford 60 •
A sample of the strip metal is rolled with back tension employed at a
constant roll speed to effect an appropriate fractional reduction in thickness
and the roll force and roll torque measured continuously. The back tension
is gradually increased until the neutral plane coincides with the exit plane
258 ENGINEERING PLASTICITY

of the roll gap. The sense of the frictional stress then becomes unidirectional,
that is, from the entrance to the exit side of the roll gap.
Equations (7.167) and (7.168) for the roll force and roll torque, respectively,
then reduce to

(7.170)
and

Hence (7.171)
which is a dimensionless parameter that can be determined experimentally
and independent of the rolling theory.

(f) Roll flattening


A sufficiently accurate correction for roll flattening is obtained by assuming
the roll deforms elastically over the arc of contact to a radius, R', which is
greater than the radius, R, of the undeformed roll but the roll profile remains
circular. The following equation has b~.:n proposed by Hitchcock 47 :
R' = R { 1 + (cF/ Ll h)} (7.172)
where F is the roll force per unit width of strip, Llh = H - h is the reduction
in strip thickness and c is a constant equal to 3.34 x 10- 4 in 2 jtonf =
5.15 x 10- 3 mm 2 /N for steel rolls.
(g) Angle of bite and maximum draft
For normal industrial rolling practice, the neutral plane is within the arc
of contact. If the neutral plane is not within the arc of contact, the frictional
resistance at roll surfaces acts in one direction and the rolls may slip on the
strip.
If the strip is required to enter the rolls unaided then the component
of the roll surface friction in the direction of rolling must exceed that of the
component of radial pressure at the entry plane. Therefore 11(pdA) cos</> 1 >
(pdA) sin </> 1 where A denotes area of contact between the rolls and the strip.
Hence 11 > tan </> 1 and the maximum value of
</> 1 =tan- 1 11 (7.173)
is usually known as the angle of bite.
Let L be the distance from the entry plane to the exit plane measured in
the direction of rolling. By referring to figure 7.42(a) it will be seen that
(R') 2 = I3+ {(R'f- (Llh/2) }2
~ I3 + (R') 2 - R' Llh

Hence (7.174)
PLANE STRAIN PLASTIC DEFORMATION 259
Also tan ¢ 1 = L/ { R'- (~h/2)} ;l::j (~h/R') 1 1 2

Since J.1 ~tan ¢ 1 and J.1 ~ (~h/R') 112

The reduction in thickness of the strip H - h = ~h is sometimes referred


to as the draft. The maximum draft is thus given by
(7.175)
It follows that large diameter rolls and high frictional resistance are
consistent with achieving maximum reduction in thickness or draft. However,
it is possible to roll with a greater draft than that given by equation (7.175)
if the strip can be moved into the roll gap by external pressure.

REFERENCES
1. Hill, R., The Mathematical Theory of Plasticity, ch. VI and app. III,
Oxford University Press, London (1950)
2. Johnson, W., Sowerby, R. and Haddow, J. B., Plane Strain Slip Line
Fields: Theory and Bibliography, ch. 3, and app. 2, Edward Arnold,
London (1970)
3. Hencky, H. Z., Uber einige statisch bestimmte Faile des Gleichgewichts
in plastischen Korpern, Z. angew. Math. Mech., 3, 241 (1923)
4. Christopherson, D. G., Oxley, P. L. B. and Palmer, W. B., Orthogonal
cutting of a work-hardeningmaterial, Engineering, Lond., 186, 113 (1958)
5. Palmer, W. B. and Oxley, P. L. B., The mechanics of orthogonal machin-
ing, Proc. lnstn mech. Engrs, 173, 623 (1959)
6. Geiringer, H., Beitrag zum Vollstiindigen ebenen Plastizitiits-problem,
Proc. 3rd Intern. Congr. appl. Mech., 2, 185 (1930)
7. Prager, W., A geometrical discussion of the slip line field in plane
plastic flow, Trans. R. Inst. Techno/., Stockholm, 65, 27 (1953)
8. Prager, W. and Hodge, P. G., Theory of Perfectly Plastic Solids, Wiley,
New York (1951)
9. Thomas, T. Y., Plastic Flow and Fracture in Solids, Academic Press,
London (1961)
10. Prager, W ., The theory of plasticity: A survey of recent achievements,
Proc. Instn mech. Engrs, 169, 1 (1955)
11. Ford, H., The theory of plasticity in relation to engineering application.
J. appl. Math. Phys., 5, 1 (1954)
12. Alexander, J. M., Deformation modes in metal forming processes,
Proc. Conf. Technol. Engng Manuf., Pap. 42, lnstn Mech. Engrs (1958)
13. Bishop, J. F. W., Green, A. P. and Hill, R., A note on the deformable
region in a rigid-plastic body, J. Mech. Phys. Solids., 4, 256 (1956)
14. Green, A. P., Plastic yielding of notched bars due to bending, Q. J.
Mech. appl. Math., 6, 223 (1953)
260 ENGINEERING PLASTICITY

15. Green, A. P., On the use of hodographs in problems of plane plastic


strain, J. Mech. Phys. Solids., 2, 73 (1954)
16. Ford, H., Advanced Mechanics of Materials, pt IV, ch. 32, p. 518,
Longmans Green, London (1963)
17. Hill, R., Discussion of a paper by E. Siebel, Application to shaping
processes of Hencky's laws of equilibrium, J. Iron Steel Inst., 156,
513 (1947)
18. Hill, R., A theoretical analysis of stresses and strains in extrusion and
piercing, J. Iron Steel Inst., 159, 177 (1948)
19. Prandtl, L. Uber die Harte Plastischer Korper, Nachr. Ges. Wiss.
Gottingen, p. 74 (1920)
20. Bishop, J. F. W., On the complete solution to problems of deformation
of a rigid-plastic material, J. Mech. Phys. Solids, 2, 43 (1953)
21. Hill, R., On the mechanics of cutting metal with knife-edge tools, J.
Mech. Phys. Solids, 1, 265 (1953)
22. Hill, R., The Mathematical Theory of Plasticity, ch. IX, p. 254, Oxford
University Press, London (1950)
23. Hill, R., Lee, E. H. and Tupper, S. J., A theory of wedge indentation of
ductile materials, Proc. R. Soc., Ser. A, 188, 273 (1947)
24. Grunzweig, J., Longman, I. M. and Petch, N. J., Calculations and
measurements on wedge-indentation, J. Mech. Phys. Solids, 2, 81 (1954)
25. Johnson, W. and Kudo, H., The cutting of metal strips between partly
rough knife edge tools, Int. J. mech. Sci., 2, 294 (1961)
26. Johnson, W., The cutting of round wire with knife-edge and flat edge
tools, Appl. Sci. Res. (A), 7, 65 (1957)
27. Mahtab, F. U. and Johnson, W., The cutting of strip with knife-edge and
flat-edge face tools, J. Mach. Tool Des. Res., 2, 335 (1962)
28. Green, A. P., A theoretical investigation of the compression of a ductile
material between smooth flat dies, Phil. Mag., 42,900 (1951)
29. Collins, I. F., Geometric properties of some slip line fields for compres-
sion and extrusion, J. Mech. Phys. Solids, 16, 137 (1968)
30. Johnson, W. and McShane, I. E., A note on calculations concerning the
plastic compression of thin material between smooth plates under
conditions of plane strain, Appl. Sci. Res., A, 9, 169 (1960)
31. Prandtl, L., On the penetration hardness of plastic materials and the
hardness of indenters, Z. angew. Math. Mech., 1, 15 (1920)
32. Hill, R., Lee, E. H. and Tupper, S. J., A method of numerical analysis of
plastic flow in plane strain and its application to the compression of a
ductile material between rough plates, Trans. Am. Soc. Mech. Engrs,
J. Appl. Mech., 18, 46 (1951)
33. Alexander, J. M., The effect of Coulomb friction in the plane strain
compression of a plastic-rigid material, J. Mech. Phys. Solids, 3, 233
(1955)
34. Sachs, G., Spanlose Formung der Metalle, Springer Verlag, Berlin (1931)
PLANE STRAIN PLASTIC DEFORMATION 261
35. Hoffman, 0. and Sachs, G., Introduction to the Theory of Plasticity for
Engineers, McGraw-Hill, New York (1953)
36. Sendzimir, M. G., The Sendzimir cold strip mill, J. Metals, N. Y., 8,
1154 (1956)
37. Underwood, L. R., The Rolling of Metals, John Wiley, New York (1950)
38. Larke, E. C., The Rolling of Strip, Sheet and Plate, Chapman and Hall,
London (1957)
39. Alexander, J. M. and Brewer, R. C., Manufacturing Properties of
Materials, ch. 5, p. 256, Van Nostrand Reinhold, London (1963)
40. Avitzur, B., Metal Forming: Processes and Analysis, ch. 15, p. 436,
McGraw-Hill, New York (1968)
41. Rowe, G. W., An Introduction to the Principles of Metalworking, ch. 9,
p. 194, Edward Arnold, London (1965)
42. Tarnovskii, I. Y., Pozdeyev, A. A. and Lyashkov, V. B., Deformation of
Metals During Rolling (English translation), Pergamon Press, Oxford
(1965)
43. Tselikov, A. I. and Smirnov, V. V., Rolling Mills (English translation),
Pergamon Press, Oxford (1965)
44. von Karman, T., Beitrag zur Theorie des Walzvorganges, Z. angew,
Math. Mech., 5, 139 (1925)
45. Orowan, E., The calculation of roll pressure in hot and cold flat rolling,
Proc. Instn mech. Engrs, 150, 140 (1943)
46. Bland, D. R. and Ford, H., The calculation of roll force and torque in
cold strip rolling with tensions, Proc. Instn mech. Engrs, 159, 144 (1948)
47. Hitchcock, J., Roll neck bearing, Report of Am. Soc. Mech. Engrs,
Special Research Committee on Heavy-duty Anti-friction Bearings
(1935)
48. Stone, M.D., Rolling of thin strips, Iron Steel Engr, 30(2), 61 (1953)
49. Hill, R., The Mathematical Theory of Plasticity, ch. VII, p. 188, Oxford
University Press, London (1950)
50. Ford, H., Roll hard materials in thin gauges: Basic considerations, J.
Inst. Metals, 88, 193 (1960)
51. Avitzur, B., Maximum reduction in cold strip rolling, Proc. Instn mech.
Engrs, 174, 865 (1960)
52. Alexander, J. M., A slip-line field for the hot rolling process, Proc. Instn
mech. Engrs, 169, 1021 (1955)
53. Alexander, J. M. and Ford, H., Limit analysis of hot rolling, in Progress
in Applied Mechanics: The Prager Anniversary Volume, p. 191, Collier-
Macmillan, London (1963)
54. Druyanov, A. B., Kinematic problems in the theory of the plane plastic
flow of ideally plastic bodies, in Investigation of Processes of Plastic
Deformation of Metals, p. 134, Moscow (1965)
55. Firbank, T. C. and Lancaster, P, R., A suggested slip line field for cold
rolling with slipping friction, Int. J. mech. Sci., 7, 847 (1965)
262 ENGINEERING PLASTICITY

56. Johnson, W. and Kudo, H., The use of upper bound solutions for the
determination of temperature distributions in fast hot rolling, Int. J.
mech. Sci., 1, 175 (1960)
57. Ford, H., Researches into the deformation of metals by cold rolling,
Proc. Instn mech. Engrs, 159, 115 (1948)
58. Watts, A. B. and Ford, H., On the basic yield stress curve for a metal,
Proc. Instn mech. Engrs, 169, 1141 (1955)
59. Hessenberg, W. C. F. and Sims, R. B., The effect of tension on torque
and roll force in cold strip rolling, J. Iron Steel Inst., 168, 155 (1951)
60. Whitton, P. W. and Ford, H., Surface friction and lubrication in cold
strip rolling, Proc. Instn mech. Engrs, 169, 123 (1955)
61. Wilcox, R. J. and Whitton, P. W., The rolling of thin titanium strip,
J. Inst. Metals, 88, 200 (1960)
62. Bedi, D. S. and Hillier, M. J., Hydrodynamic model for cold strip rolling,
Proc. Instn mech. Engrs, 182(1), 153 (1967)
8
Plastic Strain with Axial
Symmetry

The preceding chapter considered the application of plasticity theory to


the quasi-static, plane strain deformation of essentially an isotropic, rigid-
perfectly plastic material. For some applications, allowance was made for
the strain-hardening of engineering metals by means of a technological
approach. The differential equations for force equilibrium and for strain-rate
are hyperbolic. This enables the method of the slip line field technique to be
used for the solution of a number of problems and the knowledge required
to solve these problems is now well developed.
However, many components are produced by the manufacturing processes
of forging, drawing, extrusion and other related processes which have an
axisymmetric form. Such bodies are generated by revolution about an axis
of symmetry which can be arranged to coincide with the axis of the cylindrical
coordinate system (r, lJ, z). The mathematical analyses of axisymmetric
problems are, in general, difficult even when the conditions are idealised
since, if the von Mises yield criterion and the associated flow rule are used,
the governing equations are not hyperbolic 1 . 'One-dimensional' axisym-
metric problems in which the stress-strain distribution depends on one
independent variable, say the radius r, only are relatively simple to solve
although these may sometimes require the application of numerical methods.
When, however, the Tresca yield criterion and associated flow rule are
employed then, depending upon the position of the stress point on the yield
locus, the governing equations are kinematically or statically determinate
as shown by Shield2 • The papers by Hill 3 and Richmond and Morrison4
are concerned with ideal forming processes for isotropic, perfectly plastic
materials and the latter authors have determined ideal die profiles for
axisymmetric plastic flow.

8.1 FUNDAMENTAL EQUATIONS FOR PLASTIC STRAIN


WITH AXIAL SYMMETRY
The fundamental equations relating to axisymmetric stress and strain-rate
264 ENGINEERING PLASTICITY

... i>'Tzr £
.----- 'zr+- .uz
bz

Figure 8.1. Stresses acting on an element in a region of plastic strain with axial
symm~try

are most conveniently expressed in terms of cylindrical coordinates (r, 0, z),


the z axis being the axis of symmetry.
All the displacements are independent of the polar angle 0 since the flow
is confined to the meridian planes and if the body is not subjected to torsion
the azimuth velocity component v = 0 and the shear stress components
-r;rB = <z11 = 0. The non~vanishing stress components are thus ur, u 11 , uz and
•rz and the non-vanishing components of velocity are u and w referred to the
r and z directions, respectively. The stresses acting on an element are as

l
shown in figure 8.1.
If the inertia forces are insignificant and other body forces are absent, the
differential equations of force equilibrium (2.10) reduce to

O(J'r
a,: 07:zr (J'r- (}'8
+ oz + - r - =
0
(8.1)

a.zr + 'tzr + 0(J'z =0


or r oz
The strain-rate components are given by
er = oujor 88 = ujr ~z = iJ~/oz } (8.2)
Yzr = (oujoz) + (owjor) Yzll = Y11r = 0
and t)le condition of incompressibility is

er + ell + liz : 0 } (8.3)


or E:r + E:o + E:z- 0
PLASTIC STRAIN WITH AXIAL SYMMETRY 265
The equivalent strain-rate, i, for plastic flow with axial symmetry is
i = (2/3)lf2[UY + Uo)2 + (sz)2 + {(rz.?/2} ]112 (8.4)
The von Mises yield criterion of equation (4.10) expressed in cylindrical
coordinates becomes
(<rr- 11o) 2 + (11 6 - azf + (<rz- aY + 61:;r = 6k 2 (8.5)
and the Levy-von Mises stress-strain rate relations of equations (5.37)
expressed in cylindrical coordinates can be stated as
srf(ar + p) = io/(ao + p) = szf(az + p) = Yzr12•zr =A.> 0 (8.6)
where p is the hydrostatic pressure defined by
(8.7)
Hence
i;r = oujor = ~(2<rr- O"o- O"z) )
e6 =
ujr = i(2a6 - O"z- <rr)
(8.8)
i;z = owjoz = A.(2<rz- O"r- <ro)
2yzr = (oujoz) + (owjor) = 6i•zr
It is necessary that 116 = <rr when r = 0 if the stresses are not to be infinite.
Since the axis is a principal stress direction •zr
= 0 and <rr- <rz = ± (3) 112 k =
± Y when r = 0. Also the velocity component u = 0 since the axis is a
streamline. Equations (8.1), (8.3) and (8.8) are a system of seven equations
for seven unknowns which include four stress components, two velocity
components and i. However, there are only three equations involving the
stresses alone. A fourth equation for stresses can be derived from equations
(8.8) by eliminating u, w and i. In an actual problem the stress boundary
values may be known but not their derivatives. It will therefore be necessary
to resort to the velocity equations and the problem is thus not statically
determinate. For this reason a separate analysis of the stress and velocity
fields is not possible.
Hill has derived certain relations along lines of maximum shear stress
or slip lines on the meridian plane. If the maximum shear stress is denoted
by 1: and the mean compressive stress - (<rr + <rz)/2 in this plane by p',
the principal stresses are then p' ± 1: and 116 , respectively. The von Mises
yield criterion of equation (8.5) becomes
'1:2 + {(<ro + p'f/3} = kz = Y2j3}
(8.9)
or 3<2 + (<ro + p')2 = 3kz = y2
The stress equations along the slip lines become
dp' + 2<dc/> + (a 8 + p'- • cot cf>)(drjr) = (o<josp)ds" along an a line}
and dp'- 2<dc/> + (a 6 + p'- 1: tan cf>)(dr/r) = (o</os")dsp along a f3line
(8.10)
266 ENGINEERING PLASTICITY

The IX and f3 lines are distinguished by the sense of the shear stress, 't', using
the same convention as for the plane strain theory, ojosrz and iJjiJsp are
the space derivatives along the O£ and f3 lines, respectively, and rjJ is the angle
measured away from the r direction towards the positive z direction which
implies that rjJ is the angle that the tangent to an O£ line makes when measured
anticlockwise from the z axis.
The velocity compatibility equations corresponding with the Geiringer
equations (7.41) for plane strain are
du- vdrjJ + (u + v cot r/J)(dr /2r) = 0 along an IX line }
dv + udrjJ + (v + u tan rjJ)(dr/2r) = 0 along a f3line (8·11)
These equations differ from the Geiringer equations only in the extra
terms in r which vanish as r -+ oo. It should be noted that u and v are the
velocity components in the O£ and f3 directions, respectively.
If the Tresca yield criterion is adopted instead of the von Mises yield
criterion of equation (8.9) then 't' = k = Y /2 and if G 0 is assumed to be the
intermediate principal stress, G 2 , equations (8.10) become

dp' + 2kdrjJ + (G0 + p'- k cot rjJ)(dr/r) = 0 along an O£ line (8.12)


dp'- 2kdrjJ + (G0 + p'- k tan rjJ)(dr/r) = 0 along a f3line
Equations (8.12) differ from the Hencky stress equations for plane strain
deformation only in the terms in r and as r -+ oo reduce to the Hencky
stress equations (7.33). However, Hill has pointed out that this does not
mean the slip lines are characteristics.
To obtain solutions of these equations, Haar and von Karman 5 introduced
the assumption that G 6 is equal to either of the other principal stresses
in which case G 6 + p = ± k and then equations (8.12) become hyperbolic.
With the stress field obtained, no velocity distribution can generally be
determined which also satisfies the Levy-von Mises relations of equations
(8.6) since there are three independent equations and only two unknown
velocity components. Calculations based on the assumption therefore
introduce an error of unknown magnitude.
A slip line field solution for the smooth, axisymmetric flat punch indenta-
tion of a semi-infinite m:edium involving the above assumption and the
Tresca yield criterion with its associated flow rule was obtained by Shield 2 •
He suggests that his solution is a complete solution for the assumed material
and also obtained the associated velocity field.

8.2 APPROXIMATE SOLUTION FOR THE QUASI-STATIC


AXISYMMETRIC UPSETTING OF A CIRCULAR CYLINDER
An approximate equation was derived by Siebel 6 in 1923 for the mean
platen pressure during the quasi-static compression of a short circular
cylinder of a rigid-perfectly plastic material having a height to diameter
PLASTIC STRAIN WITH AXIAL SYMMETRY 267
Pressure distribution (solid of revolution)

,(>•''-"'"-r
•Friction hill'

..v..
y
JY exp(2 ~a/h)

I Platen I L..
!
(a)
I
I
h

I
L 1- 2a1 Dia.

r
..1 J

(b) h

Figure 8.2. Quasi-static compression of a circular cylinder: (a) physical diagram


and 'friction hill'; (b) stresses acting on an element of the cylinder at any radius r

ratio of about unity. It is assumed that Coulomb friction is present at the


cylinder-platen interfaces such that the coefficient of friction, Jl, is low
and that barrelling of the cylinder is insignificant. The three normal stresses
are (Jr' (J 6 and (Jz which are assumed to be principal stresses and independent
of z. The diameter and height of the cylinder at any instant are 2a and h,
respectively, shown in figure 8.2(a). The stresses acting on an element of
the cylinder are shown in figure 8.2(b). Shear stresses are neglected other
than the frictional stresses 'zr = Jl(Jz = JlP at the interfaces between the
end faces of the cylinder and the platens. The normal platen pressure at
any radius r is (Jz = p.

8.2.1 Normal platen pressure distribution and the 'friction hill'


Resolving forces exerted on the element in a radial direction produces
268 ENGINEERING PLASTICITY

(ar + dar)(r + dr)hdO- arrhde- 2pprd0dr- 2a6 sin(d0/2)hdr = 0 (8.13)


for the condition of equilibrium if inertia forces are insignificant and other
body forces are absent. Dividing by hdO and neglecting second order quanti-
ties equation (8.13) simplifies to rdar + ardr- (2J1prdrjh)- a6 dr = 0. If it is
now assumed that the radial and circumferential stresses are equal, that
is, ao = ar
dar- (2J1pdr/h) = 0
or darfdr = 2J1p/h (8.14)
The assumption that a 6 = ar can be partially justified since from the
Levy-von Mises stress-incremental strain relations expressed in cylindrical
coordinates
dar/ a~ = da6 / a~ = dazl a~
Also, the condition of incompressibility is that nr 2 h =a constant. Hence

2drjr + dhjh = 0
or 2da6 + daz = 0
However dar + da6 + daz = 0
Hence da6 =dar
and

Assuming the Tresca yield criterion to be applicable, ar- (- p) = Y where


Y is the uniaxial yield stress in compression. Therefore
dar+dp=O
or dar=- dp
Hence dpjdr = - 2J1p/h (8.15)
Transposing equation (8.15) and integrating In p = - (2w/h) +A where
A is a constant of integration. When r = a, ar = 0 and p = Y. Therefore
In Y = - (2J1a/h) +A. Substituting for the constant of integration A
ln(p/Y) = 2Jl(a- r)/h. Therefore
pjY = exp{2J1(a- r)/h} (8.16)
When r = 0 at the geometric axis of the cylinder, the normal platen pressure
is a maximum. Hence

Pmax = Y exp(2J1a/h) (8.17)


Equation (8.16) shows that the normal platen pressure varies with the
radius, r, across the cylinder in an exponential manner. This pressure
distribution or 'friction hill' is illustrated in figure 8.2(a).
PLASTIC STRAIN WITH AXIAL SYMMETRY 269
If the coefficient of friction, Jl., is low then the exponential term in equation
(8.16) can be expanded as a series and truncated after the second term to give
p/Y ~ 1+ {2JJ.(a- r)/h} (8.18)

8.2.2 Force exerted by the platens and mean platen pressure


The total compressive axial force, F, exerted by the platens on the cylinder is

F= f>2nrdr

~ 2nY f:[1 + { 2JJ.(a- r)/h} }dr


Hence F ~ na 2 Y { 1 + (2JJ.a/3h)} (8.19)
and the mean platen pressure, p = F jna 2 is
p ~ Y {1 + (2JJ.a/3h)} (8.20)
If it assumed that the height to aiameter ratio of the cylinder is unity,
that is, ajh = 1/2 and the axial force, F, is not to be increased by more
than 1 per cent, then the coefficient of friction, J1. < 0.03, indicating how
efficient the lubrication must be at the cylinder-platen interfaces.
The compression of a circular cylinder has been further considered by
Shield 2 and Bishop 7 and the forging of thin discs was examined in a similar
manner by Schroeder and Webster 8 • The compression of short circular
hollow cylinders has been analysed by Avitzilr 9 and also by Hawkyard
and Johnson 10.

8.2.3 Energy dissipation


The energy dissipated in reducing the height of the cylinder by an infini-
tesimal amount 8h is 8E = - F 8h.
The total energy dissipated in reducing the height of the cylinder from
H 0 initially to a final height H 1 is

E=- f H, Fdh= fHo Fdh


Ho H,

Substituting for the axial force, F, from the approximate equation (8.19)
where A= na 2 the cross-sectional area of the cylinder at any instant

E~ f Ho
Ht
Y A { 1 + (2JJ.a/3h)} dh (8.21)

However, the volume of the cylinder at any instant is V = Ah or A = V /h.


270 ENGINEERING PLASTICITY

Also V = na 2 h and therefore a= ( V jnh )1 12 . Equation (8.21) can be rewritten


as

~ Yln (H 0 /H 1 ) - (4,uV 112 Y j9n 112 ){ ( 1/H~/ 2 )- ( 1/Hf/ 2 )}


~ Y[ln (H 0 /H 1 ) - ( 4,uV 112 j9n 112 H~i 2 ){ 1- (H 0 /H 1 ) 3 12 } J
The volume of the cylinder is also given by V = nD~H 0 /4 where D0 and
H 0 are the initial diameter and height, respectively. Therefore V 112 =
n 112 D0 H61 2 j2 and
E/V ~ Y[ln (H 0 /H 1 ) + { 2,uj9(H 0 /D 0 ) }{ (H 0 /H 1 ) 312 - 1}]
The fractional reduction in height of the cylinder is
R = (H0 - H 1 )/H0 = 1- (HtfH0 )
Therefore HtfH 0 =1-R
or H 0 /H 1 = 1/(1- R)
The energy dissipated per unit volume in reducing the height of the cylinder
from H 0 to H 1 and effecting a fractional reduction in height, R, is thus given
by
E/V ~ Y[ In { 1/(1- R)} + { 2,u/9(H0 /D 0 ) }{ 1/(1 - R) 312 - 1}] (8.22)
It will be appreciated that the approximate equation (8.22) indicates that
the energy dissipated per unit volume during the slow compression of a
cylindrical billet depends on the extent of deformation characterised by
the fractional reduction in height, R, the nature of the frictional conditions
at the billet-platen interfaces defined by the coefficient of friction, ,u, and
the initial height to diameter ratio, H 0 jD 0 • The first term on the right-hand
side of equation (8.22) is the energy dissipated per unit volume to produce
the homogeneous compression whilst the second term is that to overcome
the frictional resistance at the billet-platen interfaces.
The above analysis was developed on the assumption that the compression
was homogeneous despite the presence of Coulomb friction at the billet-
platen interfaces and that barrelling of the cylindrical surface of the billet
does not occur. This is an oversimplification and may lead to significant
error in estimating the energy dissipation if the fractional reduction in
height and the coefficient of friction are high. Material adjacent to the
platens may not move radially outwards in an unrestricted manner and
slide relative to the platens. Cones of dead metal may be formed which
PLASTIC STRAIN WITH AXIAL SYMMETRY 271
adhere to the platens and, as the platens approach each other, barrelling
occurs as previously described in section 6.3.1 and illustrated in figure 6.7(b).

8.3 FAST HOMOGEN EOUS COMPRES SION OF A CIRCULAR


CYLINDER
If a short circular cylinder is subjected to a fast compression between rigid,
parallel platens then inertia forces will be present and the plastic yielding of
the cylinder is affected. The force, energy and power requirements are
therefore different from those which may be predicted by the quasi-static
analyses of section 8.2.
For platen speeds in excess of about 300 mjs, elastic and plastic stress
waves may be propagated which are able to travel up and down the cylinder

Moving

n
I
I fr -1-
(a) h

I U -f-

Inertia force
prd9hdrfr

(b)

Figure 8.3. (a) Fast compression of a cylindrical billet between smooth, rigid,
parallel platens. The lower platen is stationary and the upper platen descends with
velocity V; (b) forces acting on an element of the billet at any radius r
272 ENGINEERING PLASTICITY

several times during the compression process. It can therefore be anticipated


that the magnitudes of the axial force which may be detected at the upper
and lower platens will be different. During the early stages of a fast compres-
sion, the upper platen is subjected to an axial force greater than that required
for quasi-static compression whilst the deforming material accelerates to
achieve the imposed velocity conditions. In the later stage of the compression,
the force exerted on the upper platen reduces as the required velocity of
compression is approached. This sequence of events and the consequences
tend to reverse for the case of the lower platen.
To assess the inertia effect during the fast compression of a cylindrical
billet it will be assumed that frictional resistance at the billet-platen inter-
faces is absent and the deformation is homogeneous. Furthermore, the
effects of stress wave propagation are considered to be insignificant implying
that the speed of compression is less than, say, 300 mls in the range 10-300
mls, which is of greatest industrial importance.
The fast compression of a cylindrical billet is illustrated in figure 8.3(a)
which shows the lower platen stationary and the upper platen descending
with a velocity, V. The radial velocity component is u and the radial accelera-
tion component isfr which are assumed to be independent of z at any instant.
If the material is incompressible, nr 2 V = 2nrhu or
u = rVI2h (8.23)
The radial acceleration,;;. = duldt is given by ( v l2h) orI ot + (r12h) t3 vI ot -
( r v 12h 2 ) ohl ot or
(8.24)
since orlot= u, oVIot = V and ohlot = - V. Combining equations (8.23)
and (8.24)
fr = (rV 2 14h 2 ) + (rVI2h) + (rV 2 12h 2 )
= (3rV 214h 2 ) + (rVI2h) (8.25)
When the velocity of the upper platen is constant, V= 0 and the radial
acceleration component reduces to
(8.26)
Considering the forces exerted on the element of the cylinder at a radius, r,
shown in figure 8.3(b), then for equilibrium in a radial direction
(ur + dur)(r + dr)hd()- O"rrhd()- 2u8 hdrsin(d()l2)- prd()hdrf. = 0 (8.27)
Dividing by hd() and neglecting the second order quantity
rdur + O"rdr- u 8 dr- prdrfr = 0
Therefore ( durldr) + (O"r- u 8 )lr- (pfr) = 0 (8.28)
Assuming u8 = ur as in section 8.2.1 and substituting for fr from equation
PLASTIC STRAIN WITH AXIAL SYMMETRY 273
(8.26) then equation (8.28) reduces to durfdr = 3p V 2 r/4h 2 • Integration
yields ur = ( 3p V 2 r 2 f8h 2 ) + constant. However, when r = a at the cylindrical
surface of the billet, ur = 0. Therefore
0 = (3p V 2 a 2 j8h 2 ) + constant
and ur = 3pV 2 (r 2 - a 2 )/8h 2 (8.29)
Assuming ur and uz = p, the platen pressure, to be principal stresses and the
Tresca yield criterion to be applicable, ur- (- p) = Y where Y is the uniaxial
yield stress. Hence p = Y- ur. Therefore
p= y + { 3p V 2 ( a2 - r2 )/8h2 } (8.30)
The axial force, F, exerted on the billet by the platens is

F= J: p2nrdr

=2nY J:[1+{3pV 2 (a 2 -r2 )/8h 2 Y}]rdr

= 2nY {J: rdr + (3pV 2 /8h 2 Y) J: (a 2 - r2 )rdr}

= 2nY {(a 2 /2) + (3p V 2 a4 /16h 2 Y)- (3pV 2 a4 f32h 2 Y)}


= na 2 Y { 1 + (3pV 2 a 2 /16h 2 Y)}
Therefore
(8.31)
where 2a and h are the current diameter and height of the cylinder, respec-
tively. It should be noted that the inertia effect on the axial force required
to achieve fast compression of the cylinder is expressed by the second term
on the right-hand side of equation (8.31) which is characterised by the
dimensionless parameter, p V 2 fY.
Consider a short cylindrical billet of mild steel to be subjected to a fast
compression so that the final height to diameter ratio is 1/2, that is a/h = 1.
Assume the density of the steel p = 7.8 g/cm 3 , the current yield stress Y = 435
MPa and the constant velocity of the upper platen V = 30 mfs. The contri-
bution to the total force, F, is then only about 0.3 per cent due to inertia
effect.
The influence of billet inertia during fast forging has been studied by
Dean 11 • The fast compression of a viscoplastic circular cylinder was consider-
ed by Haddow 12 who also presented an extension of the Siebel analysis for
plastic compression of a disc between partially rough platens. The relative
importance of viscous and inertia effects was investigated. The dynamics of
the fast forging of a cylindrical billet have also been analysed by Lippmann 13
including the condition when the upper platen velocity, V, is not constant.
274 ENGINEERING PLASTICITY

8.4 BAR AND WIRE DRAWING THROUGH A CONICAL DIE


8.4.1 General considerations
Bar and wire drawing is the process of pulling a cylindrical workpiece through
a die which is essentially of conical form to produce a reduction in diameter
with a consequent increase in length. The operation may be performed with
or without back tension applied to the bar or wire at the entrance to the die.
If back tension is employed the reduction in cross-sectional area per pass
may be increased and the die life extended.
Historically, wire drawing is a craft which dates back to ancient Egypt.
The development of the standard wire gauges was empirical, being based
largely on the choice of successive reductions possible without annealing
or breaking the wire.
Bar and wire drawing is normally a cold-working process in which close
dimensional tolerances are achieved and a good surface finish can be pro-
duced. Large bars of up to 150 mm diameter or more are given a so-called
light sizing pass reducing the diameter by about 1 to 2 mm to improve the
surface finish and dimensional accuracy. Smaller sizes of circular section
bar ~.ce drawn to give fractional reductions in area per pass of up to 0.5.
Wires may be reduced with fractional reductions in area as high as 0.9 in
sequential pa~es from the annealed state before they are re-annealed. Some
wires having a finished diameter of 0.02 mm or less are drawn through a
large number of dies before reaching the final size and may be re-annealed
several times.
Larger sizes of bar are drawn through dies on heavy drawbenches, the
end of the bar being initially reduced in diameter or pointed to pass through
the die. The reduced bar is gripped by the jaws of a pulling 'dog' which is
driven by a hook engaged in an endless chain travelling along a straight
track. The modem drawbenches are hydraulically operated. Bars of, say,
25 mm diameter are produced in lengths of about 15-20 m at speeds of
about 35 m/min for ferrous metals but at higher speeds for non-ferrous
metals. However, for smaller diameter bars and wire it is uneconomic to
produce these relatively short lengths on a straight drawbench. Instead,
a large rotating capstan is used to which the pointed end of the bar is attached
and as the capstan rotates the wire is drawn through the die. In this way,
long uninterrupted coils of wire can be produced and by welding coils
together continuous manufacture is possible. Speeds of about 180 m/min
are achieved and much higher speeds up to about 2000 m/min are possible
during the drawing of the smallest diameter wires.
Mass production of the smallest sizes of wire usually involves progressive
reductions made in tandem on multi-hole wire drawing machines. Capstans
are provided between each die and the rotational speeds of the capstans are
adjusted throughout the train to accommodate the increasing speed of the
wire as its length is increased following the reduction in diameter through
the dies. After the last capstan the wire is collected on a take-up spool.
PLASTIC STRAIN WITH AXIAL SYMMETRY 275
Tungsten carbide dies are used for the manufacture of wires from hard
materials and diamond dies are used for the finest gauges. The drawing
force is affected by the die profile which, in industrial practice, is usually
trumpet shaped, having a bell-mouthed entrance leading into a conical
portion with a semi-angle ranging from 4oto 12o followed by a short divergent
section. However, to simplify the analysis of the process, the die is assumed
to be conical.
The two principal methods of wire drawing are referred to as either
'wet' or 'dry' which differ with respect to the preparation of the wire, the
lubrication and the design of the machine. In wet-drawing, which is used
for fine wires below about 1 mm diameter and also copper and copper alloy
wires, the whole equipment is submerged in a bath of liquid lubricant.
The wire is then allowed to slip on the drawing capstans. Soap powders are
used as a lubricant in dry-drawing. The powder is picked up by the wire
from a container and there is no slip between the wire and capstan. This
method is used for all metals other than copper and copper alloys. Ferrous
alloys are often precoated with soft metal or with zinc phosphate to provide
the necessary lubrication.
One of the earliest published articles on drawing is that by Smith 14 who
examined the flow of metal during the process. In 1900, Musiol 15 reported
on the theory and practice of wire drawing at that time and, in 1928, Sachs 16
analysed wire drawing by the slab method and obtained a solution for the
drawing stresses. Following this theoretical work, Siebel 1 7 published his
analysis for the drawing force which was estimated by assuming uniform
deformation energy with frictional stresses absent. The Sachs analysis was
refined by Korber and Eichinger 18 to include the effect of redundant defor-
mation on the drawing stress when the workpiece is subjected to internal
shear distortion in passing through the die. An excellent review of the whole
subject is given by Wistreich 19 who considered the various aspects of the
wire drawing process including the stresses and strains, the heat generated
and resulting changes in temperature, die and interpass cooling, surface
finish, frictional resistance and die wear, lubrication and the properties
of drawn wire. Many of the variables in drawing have also been examined
experimentally by Wistreich 20 • These include die friction, drawing force
and the die separating force by using a split-die method which was originally
suggested by MacLellan 21 .
The slip line field theory has been applied to the plane strain analogue of
bar and wire drawing by Hill and Tupper 22 • The effect of friction was investi-
gated by Green and Hil1 23 and the effect of back tension by Bishop 24 . A
review of plane strain theories of the drawing process has been given by
Green 25 •
There have been many attempts to use plasticity theory to analyse the
problem of plastic flow during bar and wire drawing. Nearly all theories
neglect strain-hardening of the workpiece material or allow for it by intro-
ducing a mean yield stress. The coefficient of friction at the workpiece-die
276 ENGINEERING PLASTICITY

interface is assumed to remain constant and there is considerable difficulty


in assessing the redundant deformation which occurs. The most rigorous
analysis of the problem is probably that due to Shield 26 . However, this
neglects strain-hardening, assumes the coefficient of friction at the workpiece-
die interface to remain constant and the stress distributions at entrance and
exit of the die are not investigated.
Wire drawing is discussed in relation to lubrication by Christopherson
and Naylor 27 and also by Thomson, Hoggart and Suiter28 • A survey of
wire-drawing theory and bibliography up to about 1969 has been given by
Johnson and Sowerby 29 •

8.4.2 Quasi-static, bar and wire drawing through a conical die with and
without back tension (Coulomb friction present at the workpiece-die
interface)
The approximate analysis presented here is essentially that following the
method of Sachs in which the die semi-angle, tX, and the coefficient of friction,
J.4 at the workpiece-die interface are assumed to be constant and the work-
piece material is rigid-perfectly plastic having a constant yield stress Y.
Lubrication is assumed to be efficient such that the coefficient of friction
is low.
Figure 8.4 shows a circular section bar or wire of initial diameter D 1 = 2R 1
which is drawn through a conical die of semi-angle, tX, to reduce its diameter
to D 2 = 2R 2 at exit from the die. The drawing stress is t, pis the normal die
pressure and s is the stress due to back tension at the entrance to the die.
The stresses are also shown which act on an elemental frustrum in the
plastically deforming region at a distance, z, from the exit of the die where
the diameter d = 2r. The length of the elemental frustrum in contact with the
die is ds.
Exit Entrance

Figure 8.4. Drawing of bar or wire through a straight conical die showing the
stresses acting on an elemental frustrum
PLASTIC STRAIN WITH AXIAL SYMMETRY 277
From the geometry of the situation
tan tX = dr/dz
ds cos tX = dz = dr cot tX
dssin tX = dr
Resolving forces exerted on the elemental frustrum in the axial direction
(uz + duz){ n(r + dr) 2 } - uznr 2 + p2nr ds sin tX + J.tp2nrdscos tX = 0 (8.32)
for equilibrium. Neglecting differential quantities of the second order,
equation (8.32) reduces to
2{ uz + p( 1 + J.t cot tX) }(dr/r) + duz = 0 (8.33)
Also for equilibrium in the radial direction
0" r2nr dz - p2nr ds cos tX + J.tp2nr ds sin tX = 0 (8.34)
where ur is the radial stress.
Equation (8.34) reduces to
ur = p(1- J.t tan et) (8.35)
However, J.t tan tX is usually small in comparison with unity for optimum
drawing conditions and may be neglected. In this case, the stress state is
cylindrical and the principal stresses are u 1 = uz(tensile) and u 2 = u 3 = ur = p
(compressive).
(a) Drawing stress
Assuming uz and pare principal stresses, the Tresca yield criterion gives
(8.36)
where k is the yield shear stress and Y is the uniaxial yield stress. Also
duz + dp = 0 or
duz = - dp (8.37)
Combining equations (8.33) and (8.36)
2 {uz + (Y- uz)(1 + J.t cot et) }(dr/r) + duz = 0
or 2 { - Buz + Y(1 +B) }(dr/r) + duz = 0 (8.38)
where B = J.t cot tX. Therefore
duz/{Buz- Y(1 +B)}= 2(drjr) (8.39)
Integrating equation (8.39)
(1/B)ln{Buz + C} = 2lnr +InA
where ln A is a constant of integration and C = - Y(1 +B). Therefore
(Bu z + C) 118 = r2 A. At entry to the die where r = R 1 , uz = s, the stress due
278 ENGINEERING PLASTICITY

to back tension. Therefore


(Bs + C) 118 =RiA
and A = (Bs + C) 11B I Ri
Hence (Buz + C) 1 ' 8 = (r/R 1) 2 (Bs + C) 1 ' 8
or (Buz +C)= (r/R 1) 28 (Bs +C)
and uz = Y { (1 + B)/B}{1- (r/R 1) 28 } + s(r/R 1) 28 (8.40)
At exit from the die where r = R 2 , uz = t, the drawing stress. Therefore
t = Y{(1 + B)/B}{1- (R 2 /R 1) 28 } + s(R 2 /R 1) 28
= Y{(1 +B)/B}{1-(D 2 /D 1) 28 } +s(D 2 /D 1) 28 (8.41)
If back tension is not employed then s = 0 and
(8.42)
Consequently, when back tension is employed the drawing stress is increased.
However, the increase is less in magnitude than the stress due to back
tension.
(b) Normal die pressure
Combining equations (8.33), (8.36) and (8.37)
2{Bp + Y}(dr/r)- dp = 0
Therefore dp/(Bp + Y) = 2dr/r
which can be integrated to produce (1I B) In (Bp + Y) = 2ln r + In E or
(Bp + Y) 1 ' 8 = r2 E where B = J1. cot a and In E is a constant of integration.
At inlet to the die where r = R 1 , uz = s. Therefore, s + p = Y, from the
Tresca yield criterion and hence p = Y - s. Therefore
{B(Y- s) + YF 18 = RiE
and E = {Y(1 +B)- Bs} 118/Ri
Hence (Bp + Y) 1i 8 = (r/R 1) 2 {Y(1 +B)- Bs} 1 ' 8
Bp + Y = Y(1 + B)(r/R 1 ) 28 - Bs(rfR 1 ) 28
Therefore p = (Y /B){ (1 + B)(r/R 1 ) 28 - 1}- s(r/R 1 ) 28
(8.43)
The normal die pressure at inlet to the die where r = R 1 is given by
p= Y-s (8.44)
and increases to a maximum value at exit from the die where r = R 2 and
is given by
PLASTIC STRAIN WITH AXIAL SYMMETRY 279
p = (Y/B){(1 + B)(R 2 /R 1 ) 2 B -1}- s(R 2 /R 1 ) 2 B
= (Y /B){ (1 + B)(D 2 /D 1 ) 2 B - 1}- s(D 2 /D 1 ) 2 B (8.45)
As for the case of quasi-static, plane strain drawing of a sheet material
through a wedge-shaped die previously discussed in section 7.5.3(c), equation
(8.45) indicates that the normal die pressure is reduced if back tension
is employed. The frictional resistance and, therefore, the die wear are reduced
which is expected to result in longer die life. However, the use of back
tension has been discussed by Wistreich 30 who suggests that in wire drawing
this does not lead to any significant reduction in the wear of dies. He is
of the opinion that the nature of the wire surface and its lubrication are
more important in relation to die wear.

(c) Determination of the maximum fractional reduction of area in a single pass


A limit occurs during the drawing of bar or wire material through a
conical die due to tensile instability leading to fracture in the drawn product.
This occurs when the drawing stress, t, becomes equal to the current yield
stress of the deforming material. The fractional reduction in cross-sectional
area of the workpiece material, R, is given by
R = {n(D~- D~)/4}/(nD~/4) = (D~- D~)/D~
Therefore (8.46)
For simplicity, assume that back tension is not employed. Equation (8.42)
is then applicable and combining with equation (8.46) becomes
t/Y = {(1 + B)/B} {1- (1- R)B} (8.47)
The maximum fractional reduction in cross-sectional area, Rmax• which
can be achieved in a single pass occurs when t/Y = 1. Hence
(8.48)
If the die semi-angle tX = so and Jl is assumed to be 0.05, B = Jl cot tX = 0.05 x
7.1154 = 0.356 then
1 = (1.356/0.356) { 1 - (1 - Rmax) 0 · 356 }
Therefore
(1- Rmax) 0 · 356 = 1-0.263 = 0.737
1- Rmax = 0.737 2 · 81 = 0.424
and Rmax = 1 - 0.424 ~ 0.58 (8.49)

8.4.3 Homogeneous drawing of bar or wire through a conical die


If the deformation during the drawing process is homogeneous then frictional
resistance is absent at the workpiece-die interface and there is no redundant
280 ENGINEERING PLASTICITY

deformation. Thus f.l = 0 ahd B = J.lCOt oc = 0 but equations (8.41) and (8.42)
are not then applicable. However, if fl = 0 is substituted into the differential
equation of equilibrium for the axial direction (8.33)
2{az + p}(dr/r) + daz = 0 (8.50)
and since p = Y- az,2Y(drjr) + daz = 0 and
daz = - 2Y(dr/r) (8.51)
If the material is rigid-perfectly plastic, equation (8.51) can be integrated
to produce az =- 2Ylnr + F where F is a constant of integration. At entry
to the die, where r = Rp az =' 0. Therefore az = 2Y ln(Rtfr) and at exit
from the die where r = R 2 , az = t so that t = 2Y ln(RtfR 2 ) or
t/Y = ln(RtfR 2 ) 2 = ln(D 1 /D 2 f (8.52)
and combining equations (8.46) and (8.52)
t/Y=ln{1/(1-R)} (8.53)
The maximum fractional reduction in cross-sectional area which can be
achieved assuming homogeneous deformation during drawing is then
given by
1 = ln{1/(1- Rmax)}
1/(1 - Rmax) = exp(1) = 2.72
and 1 - Rmax = 0.368
or Rmax ""' 0.63 (8.54)
In typical drawing operations of bar and wire it can be seen, by comparing
equations (8.49) and (8.54), that the effect of frictional resistance is not
very significant in influencing the maximum fractional reduction which
can be achieved in a single pass. It is usual in industrial practice to limit
the fractional reduction achieved in a single pass to between 0.35 and 0.45.
This is because at higher values of fractional reduction it is difficult to
maintain efficient lubrication and there is the danger of metallic 'pick-up'
from the deforming material on the die surface with resulting deterioration
in surface finish of the drawn product.

8.5 EXTRUSION THROUGH A CONICAL DIE

8.5.1 General considerations


Extrusion is the process in which the workpiece, referred to as a billet, slug
or calot, is transformed into a continuous product, usually of uniform
cross-section such as circular section rod, by forcing it under high pressure,
through a die shaped to produce the required cross-section. The initial
workpiece is in the form of a cast billet or a slug sheared from bar stock
PLASTIC STRAIN WITH AXIAL SYMMETRY 281
depending on the metal and on the desired finished product. A diversity
of sections can be produced economically by extrusion with good dimen-
sional accuracy and surface finish. Products with non-uniform sections
have also been produced by the extrusion process.
The process involves a reduction in the cross-section of the workpiece
with an increase in its length. There is generally no loss of metal except
for the unextruded portion which is discarded at the end of the process
and the subsequent machining operations required to achieve the final
dimensions of the product are reduced to a minimum. The primary extrusion
process is used to refine as cast metal billets into finished or semi-finished
wrought products and the secondary extrusion process is used to convert
these products into finished engineering components. The 'as cast' metals
are normally extruded hot, that is, above the recrystallisation temperature
at relatively low speeds with billet sizes varying from about 150 mm to
500 mm diameter in horizontal hydraulic extrusion presses having capacities
of up to about 50 MN. Slugs sheared or 'cropped' from the wrought bar
are normally extruded in crank-type presses at relatively high speeds. The
workpieces are usually small and in the majority of cases are not heated.
These secondary extrusion processes are often referred to as impact extrusions.
Collapsible tubes used, for example, to contain toothpaste are typical impact
extrusion products. However, impact extrusion has been extended to the
cold extrusion of some steels and to a variety of symmetrically shaped
components which were formerly produced by machining. Extrusion is
now often able to compete favourably with conventional machining for
the manufacture of a large number of components in terms of the required
dimensional accuracy, surface finish, improved mechanical properties, less
metal wastage and lower manufacturing costs.
There are two common methods of extrusion known as (a) forward or
direct extrusion and (b) backward, indirect or reverse extrusion. In forward
extrusion the punch and billet travel in the same direction whilst in backward
extrusion they travel in opposite directions. The distinction between the
two methods was previously discussed in section 7.4.1 in relation to plane
strain extrusion and illustrated in figure 7.24.
Extrusion is one of the youngest of the metal forming processes. Pearson
and Parkins 31 have pointed out that the principle of extrusion was apparently
first applied by J. Bramah of Sheffield who, in 1797, invented a machine
to extrude lead pipes. The principle was then extended to the covering of
cables which brought about the possibility of telegraph communication
between continents by the laying of ocean cables.
In 1931, Siebel and Fangmeier 32 performed experiments in hot extrusion
and apparently first presented an analytical method of determining the
power requirements for extrusion. In the same year, a similar study was
undertaken by Sachs and Eisbein 33 and the influence of pressure and tem-
perature during extrusion were considered by Pearson and Smythe 34 .
The analysis of the extrusion process was based on the so-called total
282 ENGINEERING PLASTICITY

strain concept until the more rigorous mathematical treatment was intro-
duced by Hill 35 in 1948 for the case of plane strain extrusion employing
slip line field theory. Because slip line field techniques were found to be
effective in understanding the extrusion process, especially in accounting
for its inhomogeneous nature, the method was extensively applied thereafter
by Prager and Hodge 36 , Green 37 , Bishop 38 and principally, Johnson 39 - 42 to
various plane strain problems.
Later, Johnson 43 - 46 and Kudo 47 - 49 developed the upper-bound
approach for the approximate analysis of the process and Kudo 50 •51 extended
its application to the axisymmetric extrusion problem. A technique of
semi-empirical analysis, based on strain distribution measurements, and
known as the visioplasticity method, has been developed by Thomsen,
Frisch and other co-workers 52 - 54 which has made contributions to a better
understanding of some complicated problems. The monograph by Johnson
and Kudo 55 is a detailed survey of the mechanics of metal extrusion up to
about 1960. Important experimental and theoretical work has been per-
formed by the Plasticity Division of the National Engineering Laboratory
in Scotland 5 6 - 58 and, more recently, the warm extrusion of steel has been
investigated 59 •
The cold extrusion of steel was first developed in Germany during the
1939-45 War, the success of which depends on phosphating the billet as
a means of entrapping the lubricant. The development of the cold extrusion
of steel is surveyed by Morgan 60 and a monograph has been produced on the
cold forging of steel by Feldman 61 . The topics of frictional resistance and
lubrication in metalworking processes are considered in detail in the book
edited by Schey 62 • However, special reference should be made here to the
Ugine-Sejournet process 63 which utilises fusible glass powder as a lubricant.
This has facilitated the hot extrusion of high strength, high temperature
alloys.

8.5.2 Phases during extrusion


Typical extrusion force-punch displacement autographic diagrams for
the slow forward and backward cold extrusion of a cylindrical slug of a
well-lubricated non-hardening material through a single circular orifice,
square-faced die are shown in figure 8.5. Examination of these diagrams
suggests that there are three principal phases during the processes which
can be distinguished as (a) the coining phase, (b) the steady state phase
and (c) the post-steady or unsteady state phase. The nature of the diagrams
is not significantly affected when a strain-hardening material is extruded
through a conical die.

(a) Coining phase


At the commencement of an extrusion, the billet initially deforms elasti-
cally in compression, yields in the vicinity of the die and then in this region
PLASTIC STRAIN WITH AXIAL SYMMETRY 283
(a) Coining (b) Steady state (c) Unsteady state
phase phase phase

/' ,, ____ i_________ _


u
...
~
0

..
c

..
.5!
::lJ
~
(ii) Backward or
)(
11.1
indirect
extrusion

End of
shady state

Punch displacement

Figure 8.5. Idealised autographic extrusion force-punch displacement diagrams


for (i) forward or direct extrusion, and (ii) backward or indirect extrusion illustrating
the phases of the processes

deforms plastically to exactly fill the die and the container which expand
elastically. At the same time there is usually a limited amount of extrusion
of relatively unstrained material. During this coining phase the extrusion
force is therefore expected to increase rapidly. Duffill and Mellor64 have
discussed the high initial extrusion pressure which occurs when extruding
through conical dies.
(b) Steady state phase
As the extrusion proceeds and the punch continues to move forward in
a forward or direct extrusion the extrusion force decreases. This occurs
because the frictional resistance at the billet-container interface decreases
as the length of undeformed billet diminishes. However, during backward
or indirect extrusion, there is no relative motion between the billet and
the container and consequently frictional resistance is absent. The extrusion
force therefore remains sensibly constant during this phase.
(c) Post-steady or unsteady state phase
A more rapid rate of decrease in the extrusion force is evident at the
end of the steady state phase. The whole of the billet now constitutes the
zone of plastic deformation and the field of strain changes as the punch
advances. When the length of billet is reduced to about half the diameter of
the extruded product, a cavity or so-called pipe is initiated on the axis
at the rear end of the billet and the final phase of unsteady state deformation
occurs. The cavity gradually increases in diameter and depth so that the
extruded product during this phase is transformed into a pipe of increasing
284 ENGINEERING PLASTICITY

internal diameter. This extrusion defect of piping occurs at some point in


the process when the velocity of the billet at the axis of symmetry exceeds
the velocity of the punch. The material in the vicinity of the axis breaks
contact with the punch and the cavity begins to form. The extrusion force is
no longer uniformly distributed over the whole of the pressure pad.
The extrusion phases were apparently first distinguished by Johnson 46
for the case of plane strain extrusion. These ideas were then later applied
by Avitzur 65 to account for the phenomena in axisymmetric extrusion and
the topic is also discussed in the book edited by Hoffmanner 66 •

8.5.3 Quasi-static, open-die, forward extrusion through a conical die with


Coulomb friction present at the workpiece-die interface
The stresses acting on an elemental frustrum in the plastically deforming
region during forward extrusion through a conical die can be regarded the
same as for the drawing of circular section bar or wire shown in figure 8.4
and discussed in section 8.4.2. The differential equation.of equilibrium (8.33)
and equation (8.35) are applicable but in forward extrusion the boundary
stresses are different.
In bar and wire drawing, the optimum die semi-angle, oc, is expected to
be small such that J1. tan oc is negligible compared with unity and u r ~ p, the
normal die pressure. However, this is not generally valid for extrusion where
the optimum die angles are greater than for drawing.

(a) Extrusion pressure


Assuming uz and ur are principal stresses, the Tresca yield criterion
gives
G'z- ( - ur) = 2k = Y (8.55)
where k is the yield shear stress, Y is the uniaxial yield stress and u r is the
radial stress. Also
duz + dur = 0 or
duz = - dur (8.56)
Combining equations (8.33), (8.35) and (8.55)
2[ uz + (Y- uz){ (1 + J1. cot oc)/(1 - J1. tan oc)} ](dr/r) + duz = 0 (8.57)
Let (1 + J1. cot oc)/(1 - J1. tan oc) = p then
2{uz + (Y- uz)P}(dr/r) + duz =0
or 2{uz(l- P) + PY}(dr/r) + duz = 0
Therefore duz/{ (p- 1)uz- PY} = 2dr/r (8.58)
It should be noted that equation (8.39) for axisymmetric drawing and
equation (8.58) for forward extrusion are applicable to dies having both
PLASTIC STRAIN WITH AXIAL SYMMETRY 285
straight and curved profiles. However, it will be assumed in equation (8.58)
that fJ and Yare constants. Hence integration of equation (8.58) yields
{1/(/J -1)}ln{(fJ -1)az + y} = 2ln r +In F
where In F is a constant of integration andy = - fJY. Therefore
{ (/J- 1)az + y}1/(P-1l = r2F
At exit from the die where r = R 2 = D2 /2, the extruded product is stress
free, that is, assuming residual stresses are absent so that az = 0. Therefore
y1f<P-1l = R~ F
and F=y 1 1<P- 1 ljR~

Hence {(fJ- 1 )az + y }1i<P- 1l = ( rjR 2 f y1i<P- 1l


or (fJ- 1)az + y = (r/R 2 f<P-1Jy
Therefore az = {y/(fJ -1)}{(r/R 2)2<P- 1l -1} (8.59)
Let P. be the extrusion pressure then at entrance to the die where r = R 1 =
Dtf2, az = Pe:
Pe = {y/(fJ -1)}{(R1/Rz)2<P-1l -1}
= Y{fJ/(fJ-1)}{1-(RtfR 2)2<P- 1l}
= Y{fJ/(fJ-1)}{1-(D 1jD 2f<P- 1l} (8.60)
If 11. is sufficiently small such that tan 11. is small and J1. tan 11. becomes negligible
compared with unity then fJ = (1 + J1. cot a)/(1 - J1. tan 11.) R> 1 + J1. cot 11. =
1 + B where B = J1. cot 11. and equation (8.60) reduces to
(8.61)
Equation (8.60) estimates the extrusion pressure required for open-die
extrusion which is usually arranged with the axis of symmetry vertical and
where the billet is relatively short to ensure that buckling does not occur.
Equations (8.41) and (8.42) for drawing of bar or wire and equation (8.60)
for open-die extrusion are expected to provide underestimates in each case,
since no allowance is made for the additional energy dissipated in producing
redundant deformation. Redundant deformation in drawing is not very
significant but in extrusion is important because the optimum die angles
are greater than for drawing. If the billet is forward extruded through a
conical die from a cylindrical container then further energy is dissipated in
overcoming frictional resistance at the billet-container interface and as the
extrusion proceeds the length of the undeformed billet reduces causing the
extrusion pressure to decrease during the steady state phase.

(b) Normal die pressure


Combining equations (8.33), (8.35), (8.55) and (8.56)
286 ENGINEERING PLASTICITY

2{ Y- p(1- f.1 tan rx) + p(1 + f.1 cotrx)}(dr/r)- dp(1- f.1 tan rx) = 0 (8.62)
since O"z = Y- ar, ar = p( 1- f.1 tan rx) and daz = -dar= - dp(1- f.1 tan rx).
Therefore
2[ Y + p{f.l(cot rx +tan rx)}] (dr/r)- dp(1- f.1 tan rx) = 0
or dp/{ Y/(1- f.1 tan rx)} + p{f.l(Cot rx +tan rx)/(1- f.1 tan rx)} = 2drjr
Let f.l( cot rx +tan rx)/( 1- f.1 tan rx) = (J and 1/( 1- f.1 tan rx) = 4J then
dp/( Y 4J + pb) = 2drjr (8.63)
Integration of equation (8.63) yields
(1/J) In (pb + Y 4J) =In r 2 +In G
or
where In G is a constant of integration.
At exit from the die where r = R 2 = D 2 /2, az = 0. Hence ar = Y or p = Y 4J.
Therefore
G = { Y 4J(1J + 1)} 1 1°/R~
and (pb + Y 4J )1 1° = (r/R 2 ) 2 { Y 4J(b + 1) }1 /o
or pb+Y4J=(r/R 2 ) 20 {Y4J(b+1)}
Therefore
p = ( Y #J){ ( b + 1)(r/R 2 f 0 - 1} (8.64)
The normal die pressure, p, increases from p = Y 4J = Y /( 1 - f.1 tan rx) at
exit from the die to a maximum value at entrance to the die where r = R 1 =
D1 /2, and is given by
P = ( Y #J){ (J + 1)(R 1 /R 2 ) 20 - 1}
(8.65)
If rx is sufficiently small such that tan rx is small, and f.1 tan rx becomes
negligible compared with unity and f.1 cot rx, then
(J = (cot rx + tan rx )/( 1 - f.1 tan rx) ~ f.1 cot rx = B
4J = 1/( 1 - f.1 tan rx) ~ 1 and equation (8.65) reduces to
p=(Y/B){(1 +B)(D 1 /D 2 ) 2 B-1} (8.66)

8.5.4 Redundant deformation during drawing and extrusion through a


conical die
An estimate of the additional stress required to produce the redundant
deformation during drawing or extrusion through a conical die can be
obtained in a similar manner to that employed for plane strain drawing
PLASTIC STRAIN WITH AXIAL SYMMETRY 287

Figure 8.6. Shear distortion of an element at radius, r, during drawing or


extrusion through a conical die

and extrusion through wedge-shaped dies discussed in section 7.5.5. Referring


to figure 8.6 the elemental annulus at a radius, r, from the axis at the entrance
to the die is dz long and dr thick. It is sheared through an angle, 0, where
angle () varies from zero at the axis to a maximum of IX at the die surface.
Assuming this variation to be linear

() = (r/RdiX (8.67)

where IX is the die semi-angle and D 1 = 2R 1 is the diameter of the die at


entrance.
Energy dissipated per unit volume in shearing the elemental annulus
(8.68)
where k is the yield shear stress of the material. Therefore the energy required
to shear the annular element dE = (kr I R 1 )1X2nrdrdz and the total energy
required to produce the redundant deformation at entrance to the die

E = (2nk1Xdz/R 1 ) I R,

0
r 2 dr

= 2nkiXdzR~/3 (8.69)
The volume of material entering the zone R:~ nR~dz. Therefore the energy
dissipated per unit volume in producing redundant deformation at entrance
to the die
R:~ 2k1X/3 R:~ Y IX/3 (8.70)
for a rigid-perfectly plastic material assuming the Tresca yield criterion to
be applicable and where Y is the uniaxial yield stress.
For a strain-hardening material, equation (8.70) becomes
288 ENGINEERING PLASTICITY

Ya/3 (8.71)
where Y is the mean yield stress.
Similar expressions can be deduced for the energy dissipated per unit
volume in producing the redundant deformation at exit from the die when
the elements are again sheared to move parallel to the axis within the drawn
product. The additional stress, ared• required to produce redundant defor-
mation during drawing or extrusion through a conical die is thus given
approximately by
O"red = 2Ya/3 (8.72)
for a rigid-perfectly plastic material, where Y is the uniaxial yield stress or
O"red = 2 Ya/3 (8.73)
for a strain-hardening material if Y is the mean yield stress.
It has been shown in section 8.4.2(c) that there is a limit to drawing of
bar or wire material through a conical die due to tensile instability which
leads to fracture in the drawn product. However, during extrusion through
a conical die there is no similar limitation to forming although there are
other limitations and defects which may occur. Extrusion as a manufacturing
process is therefore attractive economically since large reductions in cross-
sectional area can be effected in a single pass which may also be advantageous
in improving the metallurgical structure of the material. Short dies having a
large included angle are thus desirable to reduce the physical size of the dies
which are often manufactured from expensive tungsten alloy steels. It follows
that the contribution to the extrusion pressure in producing redundant
deformation during extrusion is of considerable importance and is significant.
The actual extrusion pressure required for forward, open-die extrusion
which does not involve frictional resistance at a billet -container interface,
is the pressure to produce homogeneous deformation and overcome the
frictional resistance at the billet-die interface plus the pressure required to
produce redundant plastic deformation. In these circumstances the actual
extrusion pressure required to extrude a cylindrical .billet from an initial
diameter, Dp to a diameter, D 2 , through a conical die having a semi-angle,
a, and where J.l is the coefficient of friction is given by
Pe<l> = - [ Y{P!<P -l)}{(D 1/D 2 ) 2<P-l> -1} + (2 Ya/3)] (8.74)
if the material is strain-hardening and Y is the appropriate mean yield stress.
The derivation of equations (8.72) and (8.73) to estimate the additional
stress required to produce redundant deformation follows in some respects
that for wire drawing by Siebel 67 • However, it should be noted that the
internal shear distortion is assumed to occur mainly at the entrance and
exit of the die, that redundant deformation is independent of the fractional
reduction in cross-sectional area and that the principle of superposition can
be applied to make allowance for the redundant deformation. Wilcox and
Whitton 68 •69 reviewed previous theories of drawing through conical dies
PLASTIC STRAIN WITH AXIAL SYMMETRY 289
and conclude that none of them adequately predicts redundant deformation.
They compared the redundant work observed in their experiments with
values predicted by previous theories. On the basis of this comparison they
suggest that a better correlation is obtained by
(8.75)
where IX is the die semi-angle and R is the fractional reduction in cross-
sectional area. Redundant strain and redundant work factors for the drawing
of a strain-hardening material through conical dies have also been considered
by Atkins and Caddell 70 •71 •

8.5.5 Effect of frictional resistance at the billet -container interface


The effect on the extrusion pressure of frictional resistance at the billet-
container interface when extruding from a cylindrical container may be
estimated by the slab method of analysis. It is assumed that the lubrication
is sufficiently efficient for Coulomb friction to be present at the interface
but stiction does not occur. As in the previous analyses, it is also assumed
that the frictional resistance has no influence on the uniform stress distri-
bution across the billet cross-section.
The diagrammatic arrangement of figure 8.7 shows a billet being extruded
from a cylindrical container through a conical die. Also shown are the stresses
acting on a cylindrical element which is a distance, z, from the entrance to
the die. The length of the billet in the container when the die is just filled with
material to effect the full reduction is L. The discard length is l, which corres-

Entrance to die Piping occurs Pressure pad


p

I
I
I
I
._.._.,
I

r---Pe(minl
I

-----f-1
01 fI
I
I
I
I
I
I
I
I

Figure 8.7. Stresses acting on a cylindrical element of the undeformed billet


within a container
290 ENGINEERING PLASTICITY

ponds to the residual length of the billet at which the piping defect is expected
to occur. The internal diameter of the container is D1, which corresponds to
the diameter of the die at entrance, J1 is the coefficient of friction at the billet-
container interface and p is the normal pressure at the interface.
Resolving forces exerted on the cylindrical element in the axial direction
{(az + da )trDi/4}- (a trDi/4) + j1ptrD 1dz =
2 2 0 (8.76)
Therefore daz + (4jipdz/D 1) = 0 (8.77)
Assuming az and p to be principal stresses and the Tresca yield criterion to
be applicable, then

or p = Y- az
Therefore daz = - 4j1( Y- a 2 ) dz/D 1
and da2 /(a2 - Y) = 4jidz/D 1 (8.78)
Integration of equation (8. 78) yields ln (a z - Y) = (4J1z/D 1) + A where A is
a constant of integration.
At the entrance to the die where z = 0, az = Pe(l) which is the axial pressure
at entrance to the die given by equation (8.74), that is,
( az)z=o = Pe(l) = - [ Y {/1/(/1- 1) }{ (DtfDz) 2<P-l> -1} + (2Y o:/3)]
if the workpiece material is rigid-perfectly plastic having a uniaxial yield
stress, Y. Therefore
ln {p.< 1 >- Y} =A
Hence ln [(az- Y)/{P.( 1) - Y}] = 4J1z/D 1
The maximum extrusion pressure, Pe(max)' occurs at the commencement
of extrusion from the die when the billet length in the container is the initial
length L.
Therefore ln [{pe(max)- Y} / {Pe(1) - Y}] = 4j1LjD1
and {Pe(max>- Y}/{P.< 1>- Y} = exp(4J1L/D 1 )
Hence
Pe(max) = Y + {Pe(l)- Y}exp(4J1L/D 1) (8.79)
The minimum extrusion pressure, Pe(min)' occurs at the end of the steady
state phase when z = l, the discard length of the billet. Therefore
ln [{pe(min)- Y}/{Pe(l)- Y}] = 4jiljD1
and Pe(min) = Y + {Pe(1)- Y}exp(4J1l/D 1) (8.80)
For a strain-hardening material, an estimate of the maximum and mini-
PLASTIC STRAIN WITH AXIAL SYMMETRY 291
mum extrusion pressures may be obtained by substituting the mean yield
stress, Y, for the unixial yield stress, Y, in equations (8.79) and (8.80), respec-
tively.

8.6 OPTIMUM DIE ANGLES


During drawing and extrusion through conical dies both frictional resistance
at the workpiece-die interface and redundant deformation require dissi-
pation of energy although these shear processes do not contribute usefully
to the desired deformation and form of the final product. The energy dissi-
pated in producing redundant deformation increases as the die semi-angle,
IX, increases. On the other hand, the component of the frictional resistance
in the axial direction increases with long dies having a small die semi-angle, IX.
Since the energy dissipated in overcoming frictional resistance and that
dissipated in producing redundant deformation are both superimposed on
the energy required to effect homogeneous deformation there are optimum
angles for drawing and extrusion. The optimum die angle may be defined as
that die angle which minimises the drawing stress or extrusion pressure for
a particular fractional reduction in cross-sectional area and coefficient of
friction.
10

0·9 reduction R

1>
-:::o·e

.....
Ill-

"'~0·7
Ill

.: 0-6
~
...
0
'1:10·5
Ill
Ill

"
c0-4
0
·;;;
c
~0·3
c
0·2

0•1

0 2 4 6 8 10 12 14 16
Die scmi-angle,a (degrees)

Figure 8.8. Variation of dimensionless drawing stress, t/ Y, with die semi-angle,


ex, for different constant values of fractional reduction, R, for the slow drawing of copper
wire through conical dies (after Wistreich 20 )
292 ENGINEERING PLASTICITY

The experimental data obtained by Wistreich 20 for the slow drawing of


electrolytic copper wire through conical dies are reproduced in figure 8.8.
The die semi-angles vary from about 2° to 16° and the values of fractional
reduction from about 0.05 to 0.45. The points of minimum dimensionless
drawing stress, t I Y, are connected, the points of intersection thus defining
the experimental optimum die semi-angle, llCopt, for each fractional reduction.
It will be &een that the optimum die semi-angle, llCopt, increases as the fractional
reduction increases varying from about 3° when R = 0.05 to about 9° when
R = 0.45. The energy dissipated in producing redundant deformation has
been found by Johnson 72 to be greater in axisymmetric drawing than for
plane strain drawing of a strip material.
In extrusion the frictional resistance is usually greater than for drawing.
In order to reduce the frictional resistance the die semi-angles are usually
large. In addition to reducing the physical size of the die to effect large
fractional reductions and minimising frictional resistance there are other
advantages in designing extrusion dies with large included angles. These
include, for example, reducing the radial component of the pressure which
otherwise may lead to bursting of the die and also in reducing the length of
the discard. For a typical extrusion the optimum die semi-angle, llCopP is
approximately 60° or 21lC = 120°. However, square-faced dies (2()( = 180°)
are commonly used in axisymmetric extrusion.

8.7 HYDROSTATIC EXTRUSION


The possibilities of high pressure technology were first given prominence by
Bridgman 73 ~ 75 • He demonstrated that high fluid pressure could be employed
to establish a high hydrostatic pressure component during metal forming
processes to improve the ductility of the workpiece material. A bibliography
of the considerable early work by Bridgman and also on high pressure
technology have been given by Zeitlin 76 •77 .
In the simple process of hydrostatic extrusion, the billet is forced through
a conical die by fluid at high pressure from a high pressure extrusion chamber
into a low pressure chamber. The high fluid pressure may be generated by
means of a press ram which is forced into the extrusion chamber or, alter-
natively, by using a pump with or without a pressure intensifier. The ram
type is a constant rate extrusion process, whereas the pump type produces
a constant pressure extrusion. The hydrostatic extrusion process is shown
diagrammatically in figure 8.9 which includes a ram-type extrusion chamber.
During a conventional forward extrusion, the billet is in contact with
the pressure pad at the rear end, the container, and the die at the forward
end. However, in hydrostatic extrusion the billet is surrounded by high
pressure fluid and is in contact with the die only and even contact with the
die is avoided if hydrodynamic lubrication is established. An advantage of
hydrostatic extrusion is therefore that frictional resistance at the billet-
container interface is absent enabling long billets to be extruded without
PLASTIC STRAIN WITH AXIAL SYMMETRY 293

Pressure
-£31--- relief
valve
-----!':--'+-Low pressure
fluid

Figure 8.9. Ram-type hydrostatic extrusion

the tendency to buckling or bulging. Since smaller die angles than those
used in conventional extrusion can be used a more homogeneous deforma-
tion can be effected. Because of improved ductility of the workpiece due to
higher hydrostatic pressure it has been possible to extrude brittle materials 78
and Pugh 79 has reported the successful hydrostatic extrusion of high speed
tool steel.
Fluid pressures of 3100 MPa and over have been utilised for the process
and these high pressures have presented many problems. They include, for
example, the development of suitable fluids as pressure transfer media since
normal oils which are available tend to solidify at about 3100 MPa, the
design of the high pressure chamber to withstand these pressures, the develop-
ment of seals to prevent leakage and also new methods of monitoring and
recording the high pressures.
Much design and experimental work has been devoted to the development
of hydrostatic extrusion during the last decade and considerable progress
has been made towards understanding the process and its applications.
294 ENGINEERING PLASTICITY

Important work in the field of hydrostatic extrusion has been undertaken


by the National Engineering Laboratory (N.E.L.), East Kilbride,
Scotland so- s 3 , the Metalworking Research Division of the Battelle Memorial
Institute, USA s4 and also in the USSR at the Super High Pressure Physics
Laboratory, Moscows 5 . More recent developments concerned with an
experimental high speed hydrostatic extrusion machine have been reported
by Greens 6 and Slater and Greens 7 . The design of a production ·machine
for semi-continuous hydrostatic extrusion has been presented by Lengyel
and Alexanderss. A chapter is devoted to the subject of hydrostatic extrusion
in the textbook by Avitzurs 9 which includes an analysis of the hydrostatic
extrusion of a cylindrical billet.

8.8 DRAWING AND SINKING OF A THIN-WALLED TUBE


THROUGH A CONICAL DIE

8.8.1 General considerations


Tubes of the softer non-ferrous metals are usually extruded. Also tubes are
formed from welded sheet metal and precision tubes for rockets and other
such aerospace applications can be spun with excellent dimensional accuracy.
However, the initial stage in the manufacture of seamless tubes of, for

(a) (b)

Sinking

(c) (d)

Figure 8.10. The major industrial methods of cold-drawing tubes with internal
support using (a) a moving cylindrical mandrel, (b) a fixed conical plug, (c) a floating
conical plug, and without internal support by (d) sinking
PLASTIC STRAIN WITH AXIAL SYMMETRY 295
example, steel and the high strength metals involves the elongation of a
solid cylindrical billet in a hot rotary piercing process.
In order to produce seamless tubes of the desired dimensional accuracy
and surface finish, most hot pierced tubes are redrawn one or more times in
a cold-drawing process. Tube drawing is the process of reducing the external
diameter and wall thickness of a thick-walled tube by drawing it through a
die with some form of internal support provided. There are four major
cold-drawing processes employed for the final stages in the manufacture of
seamless tubes, as illustrated diagrammatically in figure 8.10. The greatest
reductions in external diameter and wall thickness are achieved using a
moving mandrel as shown in figure 8.10(a). With this arrangement the
frictional resistance at the internal surface of the tube has the same sense
as the drawing force. The tube material is prevented from contracting circum-
ferentially during plastic deformation through the die and also subsequent
to this deformation. However, the use of mandrels is expensive; they are
subject to considerable wear and must be removed at the end of the process.
Plugs are therefore often used because they are relatively small and can be
manufactured from hard, wear-resistant materials. Plug drawing may be
subclassified into two processes depending on whether the plug is fixed at
the end of a stationary rod as shown in figure 8.1 O(b) or is a floating plug
as shown in figure 8.10(c). Except for copper and copper alloy tubes, the
reduction in cross-sectional area that can be effected by plug drawing is
usually limited by metallic 'pick-up'. This arises because of the difficulty in
ensuring efficient lubrication, especially since the temperature is likely to be
high at the interface between the plug and the inner surface of the tube. It
should be appreciated that the frictional resistance at this interface when
drawing with either a stationary plug or a floating plug has opposite sense
to that of the drawing force. The sense of the frictional resistance at the
interface between the die and the external surface of the tube is always
opposite to that of the drawing force.
In these three tube drawing processes there is some reduction in the
external diameter of the tube but the main deformation is in the reduction
of the wall thickness. If no mandrel or plug is used, the wall thickness is
then uncontrolled. The reduction in external diameter is usually accompanied
by a small increase in the wall thickness. This process is referred to as tube
sinking and is illustrated in figure 8.10(d).
One of the earliest investigations concerned with tube drawing is that
by Siebel and Weber 90 who studied the stress distribution and flow patterns.
A series of papers were published later by Sachs et al. 91 - 94 on the mechanics
of the drawing of thin-walled tubes through conical dies using both stationary
and moving mandrels. A theory for drawing through curved dies was
developed by Swift and Chung 95 •96 and an experimental study of plug
drawing was undertaken by Blazynski and Cole 97 • Other experimental and
theoretical investigations of the tube sinking process have been reported by
Moore and Wallace 98 and changes in wall thickness during tube drawing
296 ENGINEERING PLASTICITY

have been considered by Flinn 99 . Eilon and Alexander 100 have reported
an industrial investigation in terms of optimising the process variables in
tube drawing. In tube sinking the main requirements from an industrial
point of view are to predict the drawing force and the change in wall thickness
of the tube which is uncontrolled. These points were considered by Hill 101
who based his discussion on the earlier work of Swift.
In this chapter, the analysis of tube sinking and drawing is restricted to
the quasi-static drawing of a thin-walled tube through a conical die. The
effects of bending can then be neglected and the stress distribution across
the tube wall can be assumed as uniform.

8.8.2 Sinking of a thin-walled tube through a conical die


The sinking of a thin-walled tube through a conical die of semi-angle, oc,
to reduce the mean diameter from D 1 = 2R 1 to D 2 = 2R 2 is illustrated in
figure 8.1l(a). It is assumed that the wall of the tube is thin at any section

F
(a) (b)

(c)

Figure 8.11. (a) Tube sinking through a conical die; (b) stresses acting on an
element of the tube; (c) component of the radial force due to the circumferential stress,
<J0 , exerted on the element in a direction normal to the die face
PLASTIC STRAIN WITH AXIAL SYMMETRY 297
compared with the diameter so that the effect of plastic bending is negligible
and the variation in stress across the wall of the tube is insignificant. Each
part of the tube is subjected to the same deformation as it passes through
the die and the length of the tube is such that steady state deformation is
established.
At any mean radius, r, within the die the stresses acting on an element
are as shown in figure 8.1l(b) where q is the longitudinal stress, that is, the
stress parallel to the die face, u6 is the circumferential stress, pis the normal
die pressure and the frictional stress at the tube-die interface is -r = JLP where
JL is the coefficient of friction. The wall thickness at any mean radius, r, is t
and the length of the element is ds, parallel to the die face. The radial compo-
nent of the force exerted on the element due to the circumferential stress u6 ,
is 2u6 sin (d0/2) tds, which has a component normal to the die face as shown
in figure 8.1l(c).
Resolving forces exerted on the element in a direction normal to the die
face
prdOds- 2u6 sin( d0/2)t ds cos oc = 0 (8.81)
for equilibrium. Therefore prdOds- u6 dOt ds cos oc = 0 or
p = u6 t cos ocfr (8.82)
Resolving forces exerted on the element in a direction parallel to the die
face
(q + dq )( r + dr )( t + dt) dO - qrdOt + 2u6 sin (dO/2) tds sin oc + JLprdOds = 0
(8.83)
After simplifying, neglecting small quantities of the second order and dividing
by dOdr, equation (8.83) reduces to
rt(dq/dr) + qt(dr/dr) + qr(dtfdr) + u6 t + (JLprfsin oc) = 0 (8.84)
Combining equations (8.82) and (8.84) yields
rt(dqfdr) + qt(drfdr) + qr(dtfdr) + u6 t + JLU6 tcot oc = 0
or {d(qrt)/dr} + u 6 t(1 + JLCOt oc) = 0 (8.85)
which is the differential equation of equilibrium for tube sinking. If the
variation of tube wall thickness is negligible then dt/dr -+ 0 and equation
(8.85) becomes
r(dqfdr) + q + u6 ( 1 + JL cot oc) = 0
or r(dq/dr) + q + u6 (1 +B)= 0 (8.86)
where B = JL cot oc.
For a thin-walled tube, t cos ocfr will be small. The normal die pressure, p,
will therefore be small compared with the circumferential stress u6 • Hence
q(tensile) ~ p ~ u9 (compressive) or u 1 = q,u 2 = p and u 3 = u6 • If the normal
298 ENGINEERING PLASTICITY

die pressure, p, is negligible compared with the other two principal stresses
then the von Mises yield criterion gives ui - u 1 u 3 + u~ = Y 2 where Y is
the uniaxial yield stress of the rigid-perfectly plastic tube material. For the
same stress state, the Tresca yield criterion gives u 1 - u 3 = Y.
The von Mises yield criterion is known to give the better prediction for
yielding of most engineering metals. However, the Tresca yield criterion leads
to simpler mathematical application when the magnitudes and senses of
the principal stresses are known. In this analysis it is therefore proposed to
adopt the so-called modified Tresca yield criterion as a compromise which can
be stated in the form
(8.87)
where m is a constant derived by the method of least squares and has the
approximate value of 1.08. The graphical representation ofthe von Mises and
Tresca yield criteria for the stress state considered and plotted in the fourth
quadrant only is presented in figure 8.12. This illustrates the validity of the
modified Tresca yield criterion as a suitable compromise.
For a strain-hardening material having a mean yield stress, Y, the modified
Tresca yield criterion gives
_q- (- u 6 ) =mY (8.88)
or u6 =m Y-q (8.89)
Combining equations (8.86) and (8.89) produces
r(dq/dr) + q +(mY- q)(l +B)= 0
or r(dqjdr)- Bq + mY(l +B)= 0
Therefore dqj{Bq- mY(l +B)}= drjr (8.90)

Ratio q/v
0 0·2 0·4 0·6 0·8 1·0 1·2

Tresca: //
-0·2 q- 0)= y _ ___..,.,- / v
4'
"'
-04
/1'
"/
"/""
//
%~""" .
, "
2
vonM1ses:
2
Oj-OjCT3+0"3=Y
2

'/"
-1·0 Z," Modified Truca:
/ 0"- CT.= mY(m•l·08)
I 3

Figure 8.12. Graphical representation of the Tresca and von Mises yield criteria
illustrating the modified Tresca yield criterion
PLASTIC STRAIN WITH AXIAL SYMMETRY 299
Integration of equation (8.90) gives
( 1/B)ln(Bq +C)= lnr +In A
where In A is a constant of integration and C = -mY ( 1 + B). Hence
(Bq + C) 1 1B = rA. At the entrance to the die where r = R 1 = Dtf2, q 1 = 0.
Therefore
Cl/B =RlA
or A= cltu;Rl
and (Bq + C)l/B = (r/Rl )ClfB
Bq + C = (r/R 1 )BC
Therefore q = ( C/B){ (r/RdB- 1}
or q =mY {(1 + B)/B}{1- (r/R 1 )u} (8.91)
At exit from the die where r = R 2 = D2 /2, q2 the longitudinal stress parallel
to the die face is
q 2 = mY{(1 + B)/B}{1- (R 2 /R 1 )u}
=mY {(1 + B)/B}{1- (D 2 /D 1)u} (8.92)
The force, F, required in the axial direction is thus given by F ~ q 2 nD 2 t/
cos IX. That is

8.8.3 Variation of wall thickness during tube sinking


Assuming the tube material to be rigid-perfectly plastic then the Levy-von
Mises stress-strain increment relations are
de 2 = dt/t = dA.(u 2 - um) = (2/3)dA.[u 2 - {(u 1 + u 3 )/2}]
de 3 = dr/r = dA.(u 3 - um) = (2/3)dA.[ u 3 - { ( u 1 + u 2 )/2}] <8·94)
where the hydrostatic stress, um = (u 1 + u 2 + u 3 )/3 and if the tube material
is incompressible de 1 + de 2 + de 3 = 0 or
(dl/l) + ( dt/t) + ( dr/r) = 0 (8.95)
From equations (8.94)
de 2 /de 3 = (dt/t)/(dr/r) = [ u 2 - { (u 1 + u 3 )/2}]/[u 3 - { (u 1 + u 2 )/2}]
and if the normal die pressure, p = u 2 , is negligible compared with the
longitudinal stress, q = u 1 , and the circumferential stress, u6 = u 3 , then
(dt/t)/(dr/r) = (u 1 + u 3 )/(u 1 - 2u 3 )
= (q + u6 )/(q- 2u6 ) (8.96)
300 ENGINEERING PLASTICITY

For a rigid-perfectly plastic material


a 1 =q=mY{(1 +B)/B}{1-(r/Rd8 }
= mY f3 { 1 - ( r/ R 1 ) 8 } (8.97~

where B = J1. cot il( and f3 = ( 1 +B)/B. Also from the modified Tresca yield
criterion
a3=ao=a1-mY
= mY[/3{1- (r/R 1 ) 8 } -1] (8.98)
Combining equations (8.96), (8.97) and (8.98) gives
dt/t=(dr/r)[f3{1-(r/Rd8 } -1]/[2- f3{1-(r/R 1 ) 8 }] (8.99)
Integrating equation (8.99) and eliminating the constant of integration by
knowing that at entrance to the die, r = R 1 and t = t 1 , yields
In ( t/t 1 ) = {3/( B - 1)} In [ 2 ( r/R 1 ) 8 /2 - f3 { 1 - ( r/R d 8 } ] - 2ln ( r/R 1 )
(8.100)
At exit from the die where r = R 2 = D2 /2, t = t 2 • Therefore
In (t 2 /t 1 ) = {3/(B- 1)} In [2(D 2 /D 1 ) 8 /2- f3 { 1 - (D 2 /D 1 ) 8 } ] - 2ln(D 2 /D 1 )
(8.101)
The variation in the tube wall thickness ratio, t 2 /tt> with mean diameter
ratio, D2 /D 1 , for the sinking of a tube through a conical die having a semi-
angle, 0( = 15° and assuming the coefficient of friction J1. = 0.05 is represented
graphically in figure 8.13.
For the case considered it can be appreciated that there is a thickening
of the tube wall during a sinking process. The thickness of the tube at exit

-...."'- 1-15 ,.----,----,----.--,.----.- -,


)I= 0·05

..
0
a= 15°

....
~ HO
w
c
""u
:c
.. 1·05
0
•w
-'1
:::1
1-
1·0 0·9 0·8 0·7 0·6 0·5 0·4
Mean diameter ratio D2 /D1

Figure 8.13. Variation of the tube wall thickness ratio, t 2 /t 1 , with mean diameter
ratio, D2 /D 1 , for the sinking of a tube through a .:;onical die
PLASTIC STRAIN WITH AXIAL SYMMETRY 301
from the die, t 2 , can therefore be determined and may be substituted into
equation (8.93) to obtain an improved estimate of the axial force required
for sinking.

8.8.4 Plug drawing of a thin-walled tube through a conical die


Most tube drawing using a stationary plug or moving mandrel involves
some reduction in the internal diameter. This is at least necessary to allow
the plug or mandrel to be inserted before the commencement of the drawing
process. There is consequently some degree of sinking involved. However, it
is convenient to assume that the change in diameter is insignificant whilst

(a)

r+dr

dz/cos/3

(b)
Figure 8.14. (a) Geometry for the close-pass drawing of a thin-walled tube
through a conical die with a stationary conical plug; (b) assumed stresses acting on a
tapered element of the tube
302 ENGINEERING PLASTICITY

the main reduction is effected by a change in the wall thickness which is


nearly true for many industrial plug drawing operations. Hypothetical
close-pass drawing will be assumed which implies that the plug or mandrel
fits the tube bore exactly. If only the wall thickness is changed during the
drawing process then the circumferential strain is negligible. The geometry
for the process is illustrated in figure 8.14(a) and the assumed stresses acting
on a tapered element of the tube between the die and the stationary conical
plug are shown in figure 8.14(b). The normal die pressure and the normal
pressure at the interface between the plug and the internal surface of the tube
are both assumed to be equal to p. The coefficient of friction at the tube-die
interface is 11- 1 , and that at the tube-plug interface is 11- 2 , where 11- 1 =I= 11 2
generally.
Resolving the forces exerted on the tapered element in the axial direction
(crz + dcrz)(t + dt)2n(r + dr)- crzt2nr + J1. 1 p(dzjcos 1X)2nr cos IX
+ J1 2 p(dzjcos f3)2nr cos f3 + p( dzjcos 1X)2nr sin IX
- p(dzjcos f3)2nr sin f3 = 0 (8.102)
for equilibrium under steady drawing conditions. After simplification and
neglecting small higher order quantities equation (8.102) reduces to
dcrzt + CTzdt + pdz(J1. 1 + Jlz) + pdz(tan IX- tan /3) = 0 (8.103)
However dt = dz( tan IX- tan /3)
or dz = dt/( tan IX - tan /3) (8.104)
Combining equations (8.103) and (8.104)
dcrzt + crzdt + pdt[1 + {(~I- 1 + ~I- 2 )/(tan IX- tan /3)} J = 0
or [crz + p (1 + {(~I- 1 + ~I- 2 )/(tan IX- tan /3)}) ](dt/t) + dcrz = 0
Let (~I- 1 + ~I- 2 )/(tan IX- tan /3) = B*. Then
{ CTz + p(l + B*) }( dt/t) + dcrz = 0 (8.105)
It will be noted that equation (8.105) is of the same form as the differential
equation of equilibrium (7.118) for the plane strain drawing of sheet material
through a wedge-shaped die.
If it can be assumed that the lubrication is efficient so that the frictional
stresses are low, then crz and p can be considered as principal stresses. From
the Tresca yield criterion crz- (- p) = Y, where Y is the plane strain yield
stress since close-pass drawing is assumed and the circumferential strain is
negligible. Therefore
p= Y -(Tz (8.106)
Combining equations (8.105) and (8.106)
{ crz + (Y- crz)(1 + B*)}(dt/t) + dcrz = 0
PLASTIC STRAIN WITH AXIAL SYMMETRY 303
Therefore duJ{B*uz- ¥(1 + B*)} = dt/t (8.107)
Integrating equation (8.107)
(1/B*)ln {B*uz + C} =In t +In A
where C = - Y ( 1 + B*) and ln A is a constant of integration. Therefore
(B*uz + C) 118* = tA
However, at entry to the die where t = t 1 , uz = 0.
Therefore cl/B* = tl A
and A= C 118* /tl
Therefore (B*uz + C) 1' 8 * = (tft 1 )C 118 •
or (B*uz+C)=(t/t 1 rc
and uz=(C/B*){(t/t 1 r-1}
= ¥{(1 +B*)/B*}{1-(t/tlr} (8.108)
At the exit from the die where t = t 2 , uz = tP which is the axial stress required
for plug drawing. Therefore
tp= ¥{(1 +B*)/B*}{1-(t2/t 1 ) 8 *} (8.109)
The parameter, B*, is peculiar to the specific drawing process such that
if a stationary conical plug is employed B* = (Jl 1 + Jlz )/(tan oc - tan fJ). If
a stationary cylindrical mandrel is used, then fJ = 0 and B* =-= (Jl 1 + Jlz )/tan oc,
and for frictionless tube drawing with a conical plug or cylindrical mandrel,
jl 1 = Jlz = 0 and B* = 0. Substituting this value of B* in equation (8.107)

duz = - Y(dt/t)
Hence uz = - Y ln t +constant
When t = tp uz = 0
Therefore
At exit from the die where t = t 2 ,uz = th which is the axial stress required
for homogeneous drawing of a tube. Therefore
(8.110)

8.8.5 Tuf?e drawing through a conical die with a moving cylindrical mandrel
When a stationary plug is used the frictional resistances at the internal and
external surfaces of the tube both have opposite sense to that of the drawing
force. However, when a moving cylindrical mandrel is employed the sense
of the frictional resistance at the internal surface of the tube is reversed. This
arises because the tube elongates as it plastically deforms whilst the mandrel
304 ENGINEERING PLASTICITY

Figure 8.15. The frictional stresses acting on the internal and external surfaces of
a thin-walled tube during close-pass drawing through a conical die using a moving
cylindrical mandrel

remains undeformed. It follows that the sense of the frictional resistance at


the tube-mandrel interface is opposite to that at the tube-die interface as
shown in figure 8.15. The differential equation of equilibrium in the axial
direction and its solution to determine the axial stress, tm, required for tube
drawing employing a moving mandrel are the same as for conical plug
drawing except that tan f3 = 0 and the parameter B* is given by
(8.111)
It is, of course, conceivable that the frictional coefficient J1. 2 at the tube-
mandrel interface could be greater than J1. 1 at the tube-die interface. In this
event, the parameter B* becomes negative and the axial stress required for
drawing with a moving mandrel would be less than that for frictionless
drawing to effect the same reduction.

REFERENCES
I. Hill, R., The Mathematical Theory of Plasticity, ch. X, p. 262, Oxford
University Press, London (1950)
2. Shield, R. T., On the plastic flow of metals under conditions of axial
symmetry, Proc. R. Soc. (A), 233, 267 (1955)
3. Hill, R., Ideal forming operations for perfectly plastic solids, J. Mech.
Phys. Solids, 15, 223 (1967)
4. Richmond, 0. and Morrison, H. L., Streamlined wire drawing dies of
minimum length, J. Mech. Phys. Solids, 15, 195 (1967)
5. Haar, A. and von Karman, T., Zur Theorie der Spannungszustiinde in
plastischen und sandartigen Medien, Nachr. Ges. Wiss. Gottingen,
204 (1909)
PLASTIC STRAIN WITH AXIAL SYMMETRY 305
6. Siebel, E., Principles for calculating the work and energy requirements
in forging and rolling, Stahl Eisen, 43, pt II, 1295 (1923)
7. Bishop, J. F. W., On the effect of friction on compression and indenta-
tion between flat dies, J. M ech. Phys. Solids, 6, 132 (1958)
8. Schroeder, W. and Webster, D. A., Press forging thin sections, J. appl.
Mech., 16, 289 (1949)
9. Avitzur, B., Forging of hollow discs, Israel J. Tech., 2 (3), 295 (1964)
10. Hawkyard, J. B. and Johnson, W., An analysis of the changes in
geometry of a short hollow cylinder during axial compression, Int.
J. mech. Sci., 9, 163 (1966)
11. Dean, T. A., Influence of billet inertia and die friction in forging
processes-A simple energy approach, in Advances in Machine Tool
Design and Research, vol. B, p. 761, Pergamon Press, Oxford (1970)
12. Haddow, J. B., On the compression of a thin disc, Int. J. mech. Sci.,
7, 657 (1965)
13. Lippmann, H., On the dynamics of forging, in Advances in Machine
Tool Design and Research, p. 53, Pergamon Press, Oxford (1966)
14. Smith, 0., Flow of metals in the draw process, J. Franklin Inst., 122,
321 (1886); 123, 232, (1887)
15. Musiol, K., The drawing in extrusion presses in theory and practice,
Kingler's Polytech. J., 315, 428 (1900)
16. Sachs, G., Plasticity problems in metals, Trans. Faraday Soc., 24,
84 (1928)
17. Siebel, E., The plastic forming of metals, (translated by J. H.
Hitchcock), reprinted Steel, 43-48 (October 16, 1933-May 7, 1934)
18. Korber F. and Eichinger, A., Die Grundlagen der bildsamen
Verformung, Mitt. K- Wilhelm-Inst. Eisenforsch., Dusseld., 22, 57 (1940)
19. Wistreich, J. G., The fundamentals of wire drawing, Metal/. Rev., 3,
97 (1958)
20. Wistreich, J. G., Investigation of the mechanics of wire drawing,
Proc. Instn mech. Engrs, 169, 654 (1955)
21. MacLellan, G. D. S., Some friction effects in wire drawing, J. Inst.
Metals, 81, 1 (1952)
22. Hill, R. and Tupper, S. J., A new theory of the plastic deformation in
wire drawing, J. Iron Steel Inst., 159, pt 4, 353 (1948)
23. Green, A. P. and Hill, R., Calculations on the influence of friction and
die geometry in sheet drawmg, J. Mech. Phys. Solids, 1, 31 (1952)
24. Bishop, J. F. W., Calculations on sheet drawing under back tension
through a rough wedge-shaped die, J. Mech. Phys. Solids,2, 39 (1953)
25. Green, A. P., Plane strain theories of drawing, Proc. Instn mech. Engrs,
174(31), 847 (1960)
26. Shield, R. T., Plastic flow in a converging conical channel, J. Mech.
Phys. Solids, 3, 246 (1955)
27. Christopherson, D. G. and Naylor, H., Promotion of fluid lubrication
in wire drawing, Proc. Instn mech. Engrs, 169, 643 (1955)
306 ENGINEERING PLASTICITY

28. Thomson, P. F., Hoggart, J. S. and Suiter, J., Drawing copper wire with
a lubricant under externally generated pressure, J. Inst. Metals, 95,
152 (1967)
29. Johnson, W. and Sowerby, R., Wire drawing: A survey of theories,
Wire Ind., 137 and 249 (1969)
30. Wistreich, J. G., An investigation of back-pull wire drawing as an
industrial technique, J. Iron Steel Inst., 163,316 (1949)
31. Pearson, C. E. and Parkins, R. N., The Extrusion of Metals, 2nd. ed.,
pp. 2-14, Wiley, New York (1960)
32. Siebel, E. and Fangmeier, E., Researches on power consumption in
extrusion and punching of metal, Mitt. K-Wilhelm-Inst., Eisenforsch.,
Dusseld., 13, 29 (1931)
33. Sachs, G. and Eisbein, W., Power consumption and mechanism of
flow in the extrusion process, Mitt. Mater., S-16, 67 (1931)
34. Pearson, C. E. and Smythe, J. A., The influence of pressure and tem-
perature on the extrusion of metals, J. Inst. Metals, 45, 345 (1931)
35. Hill, R., A theoretical am.lysis of stresses and strains in extrusion and
piercing, J. Iron Steel Inst., 159, 177 (1948)
36. Prager, W. and Hodge, P. G., Theory of Perfectly Plastic Solids, ch. 6,
Wiley, New York (1951)
37. Green, A. P., Unsymmetrical extrusion in plane strain, J. Mech. Phys.
Solids, 3, 189 (1955)
38. Bishop, J. F. W., The theory of extrusion, Metal/. Rev.,2 (8), 361 (1957)
39. Johnson, W., Extrusion through wedge-shaped dies, pt I, J. Mech.
Phys. Solids, 3, 218 (1955)
40. Johnson, W., Extrusion through square dies of large reduction,
J. Mech. Phys. Solids, 4, 191 (1956)
41. Johnson, W., Partial sideways extrusion from a smooth container,
J. Mech. Phys. Solids, 5, 193 (1957)
42. Johnson, W., The plane strain extrusion of short slugs, J. Mech. Phys.
Solids, 5, 202 (1957)
43. Johnson, W., Estimation of upper bound loads for extrusion and
coining operations, Proc. Instn mech. Engrs, 173 (1), 61 (1959)
44. Johnson, W., Upper loads for extrusion through circular shaped dies,
Appl. scient Res., Sect. A., 1, 437 (1958)
45. Johnson, W., Cavity formation and enfolding defects in plane strain
extrusions, Appl. scient Res., Sect. A., 8, 228 (1959)
46. Johnson, W., An elementary consideration of some extrusion defects,
Appl. scient Res., Sect. A., 8, 52 (1959)
47. Kudo, H., An upper-bound approach to plane strain forging and
extrusion, I, Int. J. Mech. Sci., 1, 57 (1960)
48. Kudo, H., An upper-bound approach to plane strain forging and
extrusion, II, Int. J. Mech. Sci., 1, 229 (1960)
49. Kudo, H., An upper-bound approach to plane strain forging and
extrusion, III, Int. J. Mech. Sci., 1, 366 (1960)
PLASTIC STRAIN WITH AXIAL SYMMETRY 307
50. Kudo, H., Some analytical and experimental studies of axisymmetric
cold forging and extrusion, I, Int. J. Mech. Sci., 2, 102 (1960)
51. Kudo, H., Some analytical and experimental studies of axisymmetric
coldforgingandextrusion, II,Int. J. Mech. Sci.,3, 91 (1961)
52. Frisch, J. and Thomsen, E. G., An experimental study of metal
extrusions at various strain rates, Trans. Am. Soc. mech. Engrs, 76,
599 (1954)
53. Thomsen, E. G. and Frisch, J., Stresses and strains in cold extruding
2S-O aluminium, Trans. Am. Soc. mech. Engrs, 77, 1343 (1955)
54. Thomsen, E. G. and Frisch, J., Experimental and theoretical pressures
and velocity fields for various lead extrusions, Trans. Am. Soc. mech.
Engrs, 80 (I), 117 (1958)
55. Johnson, W. and Kudo, H., The Mechanics of Metal Extrusion,
Manchester University Press, Manchester (1962)
56. Watkins, M. T., Ashcroft, K. and McKenzie, J., Some aspects of the
cold extrusion of metals, Conf. Technol. Eng. Manuf., lnstn mech.
Engrs, London, sec. II, pap. 47 (March 1958)
57. Pugh, H. Ll. D. and Watkins, M. T., An experimental investigation of
the extrusion of metals (presented at Brighton), Instn Prod. Engrs
(1960)
58. Pugh, H. Ll. D., Watkins, M. T. and McKenzie, J., An experimental
investigation into the cold extrusion of steel. Sh. Metal Ind., 38 (408),
253 (1961)
59. Burgdorf, M., Extrusion of steel in the temperature range between 20
and 700°C, Metal Forming, 76 (March 1971)
60. Morgan, R. A. P., The cold extrusion of steel, J. Iron Steel Inst., 193,
285 (1959)
61. Feldman, H. D., Cold Forging of Steel, Hutchinson, London (1961)
62. Schey, J. A. (ed.), Metal Deformation Processes, Marcel Dekker,
New York (1970)
63. Sejournet, J., The hot extrusion of steel, Engineering, Lond., 177,
463 (1954)
64. Duffill, A. W. and Mellor, P. B., A comparison between the conven-
tional and hydrostatic methods of cold extrusion through conical dies,
Annals C.I.R.P., XVll, 97 (1969)
65. Avitzur, B., Steady and unsteady state extrusion, J. Engng Ind., Amer.
Soc. mech. Engrs, 89, 175 (1967)
66. Hoffmanner, A. L., Metal Forming: Interrelation between Theory and
Practice, Plenum Press, New York (1971)
67. Siebel, E., Der derzeitige Stand der Erkeuntnisser iiber die
mechanischen Vorgiinge bein Drahtsiehen, Stahl Eisen., 66, 171 (1947)
68. Wilcox, R. J. and Whitton, P. W., The cold extrusion of metals using
lubrication at slow speeds J. Inst. Metals, 87, 289 (1958)
69. Wilcox, R. J. and Whitton, P. W., Further experiments on the cold
extrusion of metals using lubrication at slow speeds, J. Inst. Metals, 88,
145 (1959)
308 ENGINEERING PLASTICITY

70. Atkins, A. G. and Caddell, R. M., The incorporation ofworkhardening


and redundant work in rod-drawing analysis, Int. J. mech. Sci., 10,
15 (1968)
71. Caddell, R. M. and Atkins, A. G., The influence of redundant work
when drawing rods through conical dies, Amer. Soc. mech. Engrs,
Pap. No. 67-WA/Prod-11 (1968)
72. Johnson, R. W., An experimental investigation of redundant work in
wire drawing, M.Sc. Dissert. University of Birmingham (1963)
73. Bridgman, P. W., Physics of High Pressure, International Textbooks
of Exact Sciences, G. Bell and Sons (1949)
74. Bridgman, P. W., Studies in Large Plastic Flow and Fracture, McGraw-
Hill, New York (1952)
75. Bridgman, P. W., Collected experimental papers, 7 vols, Harvard
University Press, Cambridge, Mass. (1964)
76. Zeitlin, A., Bibliography of publications ofP. W. Bridgman, presented
by Engineering Supervision Company, New York (January 1961)
77. Zeitlin, A., Annotated bibliography on high pressure technology:
A report of the Amer. Soc. mech. Engrs, Research Committee on
Pressure Technology, New York (1964)
78. Bobrowsky, A., Stack, E. A. and Austen, A., Extrusion and drawing
using high pressure hydraulics, Amer. Soc. Tool Manuf. Engrs, Pap.
SP65-33 (1964)
79. Pugh, H. Ll. D., Recent developments in cold forming, Bulleid
Memorial Lectures, Vols. IliA and IIIB, University of Nottingham
Press, Nottingham (1965)
80. Pugh, H. Ll. D. and Gunn, D. A., The cold extrusion of brittle materials
against a hydrostatic pressure, N. E. L. Rep. No. 31 (1962)
81. Pugh, H. Ll. D. and Ashcroft, K., The hydrostatic, or ramless, extrusion
of metals by fluid pressure, N.E.L. Rep. No. 32 (1962)
82. Pugh, H. Ll. D. and Green, D., The behaviour of metals under high
hydrostatic pressure, Part 11-Tensile and torsion tests, N.E.R.L.
Plasticity Rep. No. 128 (October 1956)
83. Pugh, H. Ll. D. and Low, A. H., The hydrostatic extrusion of difficult
metals, J. Inst. Metals, 93, 201 (1965)
84. Fiorentino, R. J., Sabroff, A. M. and Boulger, F. W., Hydrostatic
Extrusion at Battelle, Machinery Lloyd (European edn) (August 1963)
85. Beresnev, B.l., Vereschagin, L. F., Ryabinin, N. Yu. and Livshits, L. D.,
Some Problems of Large Plastic Deformation at High Pressures,
Pergamon Press, London (1963)
86. Green, D., An experimental high speed machine for practical exploita-
tion of hydrostatic extrusion, J. Inst. Metals, 93, 65 (1964-65)
87. Slater, H. K. and Green, D., Augmented hydrostatic extrusion of
continuous bar, Proc. Instn mech. Engrs, 182 (1967)
88. Lengyel, B. and Alexander, J. M., Design of a production machine for
semi-continuous hydrostatic extrusion, Proc. Instn mech. Engrs, 182,
207 (1967)
PLASTIC STRAIN WITH AXIAL SYMMETRY 309
89. Avitzur, B., Metal Forming: Processes and Analysis, ch. 11, McGraw-
Hill, New York (1968)
90. Siebel, E. and Weber, E., Stresses and metal flow in drawing tube,
Stahl Eisen, 55, 331 (1935)
91. Sachs, G., Lubahn, J.D. and Tracy, D.P., Drawing thin-walled tubing
with a moving mandrel through a single stationary die, Trans. Am. Soc.
mech. Engrs, J. appl. Mech., 11, 199 (1944)
92. Sachs, G. and Baldwin, W. M., Folding in tube sinking, Trans. Am.
Soc. mech. Engrs, 68, 647 (1946)
93. Sachs, G. and Baldwin, W. M., Stress analysis of tube sinking, Trans.
Am. Soc. mech. Engrs, 68, 655 (1946)
94. Espey, G. and Sachs, G., Experimentation on tube drawing with a
moving madrel, Trans. Am. Soc. mech. Engrs, J. appl. Mech., 14,
81 (1947)
95. Swift, H. W., Stresses and strains in tube drawing, Phil. Mag., Ser. 7,
11 (308), 883 (1949)
96. Chung, S. Y. and Swift, H. W., A theory of tube sinking, J. Iron Steel
Inst., 170, pt I, 29 (1952)
97. Blazynski, T. Z. and Cole, I. M., An investigation of the plug drawing
process, Proc. Instn mech. Engrs, 174 (28), 797 (1960)
98. Moore, G. G. and Wallace, J. F., Theories and experiments on tube
sinking through conical dies, Proc. Instn mech. Engrs, 182, 19 (1967-68)
99. Flinn, J. E., Parametric influence on the wall thickness changes and
the bulk strain behaviour of hollow drawn tubing, Trans. Am. Soc.
mech. Engrs, J. bas. Engng, p. 792 (December 1969)
100. Eilon, S. and Alexander, J. M., An industrial investigation of a process
for tube drawing, Engineer, Lond., 211,682 (1961)
101. Hill, R., The Mathematical Theory of Plasticity, ch. X, p. 269, Oxford
University Press, London (1950)
9
Plane Plastic Stress and Pseudo
Plane Stress

9.1 BASIC CONCEPTS AND ASSUMPTIONS


A stress state is referred to as plane stress if, for example, with reference to a
cartesian coordinate system of axes the stress components uz = Tyz = 'tzx = 0,
while the stress components ux, uY, -rxy are independent of z.
A state of plane stress is approximated when a thin lamina of thickness,
h, as shown in figure 9.1(a) is deformed under the influence of forces which
have their lines of action contained in its median plane. The stress components
uz, -ryz• -rzx are then small compared with the other components since the
faces of the lamina at z = ± h/2 are not subjected to force and the thickness
of the lamina is small compared with its transverse dimensions. The variation
of the stress components ux, uY, -rxy through the thickness of the lamina is
not very significant and it may be assumed that the values of these respective
components are the average values through the thickness of the lamina. At
the same time the stress components uz, Tyz• 'tzx are assumed to be zero. If
the thickness of the lamina, h, does not remain uniform during plastic defor-
mation, the stress state may be treated as approximately plane stress provided
the rate of change of thickness with distance measured parallel to the surface
of the lamina remains small.
It was shown by Hodge 1 •2 that the problem is statically determinate when
the thickness, h, is known and the known stress boundary conditions are
sufficient to establish fully plastic deformation. However, it is generally
necessary to consider the stress-strain rate relations associated with the
appropriate flow rule and the condition of incompressibility to provide
sufficient equations to solve for the unknown stresses.
In chapter 7 it was shown that plane strain problems can be solved by
employing the method of characteristics discussed in appendix 2. Plane stress
problems can also be solved in a similar manner. However, it is generally
recognised that their solution is intrinsically more difficult and the governing
equations for plane stress slip line fields are more difficult to manipulate than
those of plane strain. Hill 3 established the rigorous theory of plane stress slip
line fields. The stress equations are either hyperbolic, parabolic or elliptic de-
PLANE PLASTIC STRESS AND PSEUDO PLANE STRESS 311
z

cr2
Region of ellipticity

- Oj----+---""'*---rr-.1-+....-=-=--- Oj
,__ _ _ Region of
hyperbollcity

Trcsca hexagon

Figure 9.1. (a) Thin lamina of thickness, h, subjected to plane stress; (b) the von
Mises and Tresca yield criteria represented in the (a 1 , a 2 ) plane for plane stress illustrat-
ing the significance of the parameter w

pending on the position as defined on the yield surface, that is, the particular
combination of the principal stresses a 1 and a 2 , where a 3 = 0. The character-
istics are then, respectively, real and distinct, real and coincident or imaginary.
When the stress equations are hyperbolic the slip lines were shown by Hill
to be characteristics for both stress and velocity. The slip lines are then
equally inclined to one or other of the principal stress directions but are not
orthogonal.
The coincidence of stress and velocity characteristics was established by
Hill using the von Mises yield criterion and associated flow rule. However,
it has been shown by Prager 4 to be equally true when the Tresca yield
criterion and associated flow rule are employed. Hodge 2 and Hill 5 have
shown that when the stress equations are parabolic there is only one set of
characteristics which coincide with the direction of the numerically smaller
principal stress.
312 ENGINEERING PLASTICITY

9.2 EQUATIONS OF PLANE STRESS WITH VON MISES


YIELD CRITERION
In the circumstances stated in the previous section, the differential equations
of equilibrium of an element, hdxdy, of the lamina shown in figure 9.1(a) in
the absence of body forces have the form
+ { o(hrxy)joy} = 0
{ o(hax)fox}
(9.1)
{o(hrx)lox} + {o(hay)joy} =0
but if the thickness of the lamina, h, is assumed to be constant
(oaxfox) + (orxrfoy) = 0
(9.2)
(orxy/ox) + (oayjoy) = o
where ax, aY, rxy are the average stress components through the thickness
of the lamina.
In addition, if the material deforms plastically, the stress components
must satisfy a yield criterion which can be stated in the form
4> ( ax' ay' 7: xy) = f (a 1 ' a 2 ) = 0 (9.3)
where a 1 and a 2 are the principal stresses and a 3 = 0.
For plane stress, the von Mises yield criterion reduces to
ax2 - axay + ay2 + 3rxy
2
= 3k2 = y2 (9.4)
or in terms of the principal stresses becomes
ai- a 1 a 2 +a~= 3k 2 = Y2 (9.5)
where Y is the uniaxial yield stress and k = Y/3 112 is the yield shear stress
according to the von Mises yield criterion.
The Tresca yield criterion is given by
(9.6)
where amax and amin are the algebraic maximum and minimum values of
the principal stresses a 1 , a 2 , a 3 = 0. Unlike the case for plane strain, the von
Mises and Tresca yield criteria give different results in plane stress.
The locii of the von Mises and Tresca yield criteria in the (a 1' a 2 ) plane
are as shown in figure 9.1(b). Equation (9.5) is represented by an ellipse in
the (a 1' a 2 ) plane which has its major axis inclined at an angle + n/4 to the
+ a 1 direction and cutting off segments, 2k, on the a 1 and a 2 axes. The
principal stresses cannot exceed (2/3 112 ) Y = 2k in magnitude and the semi-
axes of the ellipse are, respectively, 2 112 Y and 2 112 k.
Depending on the sign of the principal stresses a 1 and a 2 , the maximum
shear stresses develop along different surface elements. If a 1 and a 2 are
of different sign then the maximum shear stress is rmax = (a 1 - a 2) /2 =
{ (ax - aY )2 + 4r;Y} 112 /2. If a 1 and a 2 have the same sign, for example, a 1
PLANE PLASTIC STRESS AND PSEUDO PLANE STRESS 313
and u 2 are both tensile and u 1 > u 2 then the maximum shear stress 'max =
u tf2. The Tresca yield criterion of equation (9.6) takes the form
u 1 -u 2 = ±2k= ±Yifu 1 u 2 ~0
(9.7)
u 1 = 2k = Y or u 2 = ± 2k = ± Y if u 1 u 2 0 ;:::

These equations are represented by a hexagon inscribed in the von Mises


ellipse in the (upu 2 ) plane shown by a dash line in figure 9.1(b). However,
only the solution to the plane stress equations assuming the von Mises yield
criterion to be applicable will be considered here.
Let the average (x,y) components of velocity throughout the thickness
of the lamina be denoted by (u, v) which are independent of z. If the elastic
components of the strain increment are negligible compared with the plastic
strain increment then the Levy-von Mises stress-strain rate relations are

where ex= au; ax, ty = ovfoy and Yxy = {( oufoy) + (av;ax)}


Equations (9.8) together with the equilibrium equations (9.2) and the von
Mises yield criterion (9.4) provide a system of five equations to solve for the
five unknowns ux, uY, rxy' u and v. The simultaneous solution of the nonlinear
stress and velocity equations presents considerable difficulty. However,
the solution to the stress equations (9.2) and (9.4) and to the velocity equations
(9.8) may be investigated in sequence. When the stresses are known, the
velocity equations become linear. It is then necessary to establish a velocity
field which is consistent with the stress field.
The von Mises yield criterion expressed in terms of the principal stresses
in equation (9.5) is satisfied if the following substitutions are made
u 1 = 2k cos {w - ( n/6)} and u 2 = 2k cos { w + (n/6)} (9.9)
where w is a function of (x, y) which specifies the point on the von Mises
ellipse in figure 9.1(b). If u 1 ;::: u 2 , the angle w is given by 0 ~ w ~nand is
related to the hydrostatic stress um = ( u 1 + u 2 )/3 such that
COS W = 3 112 um/2k (9.10)
The stress components ux, uY, rxy are related to the principal stresses by

: xy} = {((J 1 + (J 2) /2} ±. {((J 1


v
- (J 2) /2} COS 2¢ } (9.11)
Txy = {((J 1 - (J 2 ) /2} Sill 2¢
where <P is the angle subtended between the x axis and the first principal
direction.
Combining equations (9.9) and (9.11) yields

::} = k(3 112 cosw ± sinwcos 2¢)} (9.12)


rxy = k sin w sin 2¢
314 ENGINEERING PLASTICITY

In contrast to the case for plane strain deformation, it follows from equations
(9.12) that the stress components are bounded by ax~ 2k, aY ~ 2k and
'txy ~ k.
Combining the differential equations of equilibrium (9.2) with equations
(9.12) and simplifying produces the following two equations for the two
unknown functions of <P and w

(3 1; 2 sin w cos 2</J- cos w)(awjax) + 31; 2 sin w sin 2<jJ(awjay) )


- 2sinw(a<jJjay) = o
3112 sin w sin 2<jJ(aw;ax)- (3 1; 2 sin wcos 2</J + cosw)(awjay) (9.13)
+ 2sinw(a<jJjax) = 0
The type of the equations (9.13) can be determined as given in appendix 2.
Along some line x = x(s), y = y(s) let the functions <P = </J(s) and w = w(s)
be given. Then for the integral surface passing through the curve C
+ (a<jJjay)dy = d</J}
(a<jJjax)dx
(9.14)
(aw;ax)dx + (awjay)dy = dw
Along the curve C the tangent plane to the integral surface is given by
the partial derivatives a<jJjax, a<jJjay, awjax, awjay which can be determined
from equations (9.11) and (9.14). However, if the curve Cis a characteristic
of equations (9.13) then the derivatives are indeterminate along it.. The
determinant of the above equations and the appropriate numerators vanish.
Thus by equating to zero the determinant of the system, the differential
equations of the characteristics are
dyjdx = {3 1 / 2 sin w sin 2</J ± (3- 4cos 2 w) 112 } /(3 112 sin wcos 2</J- cosw)
(9.15)
Equating the numerators to zero and after simplifying and integrating, the
relations between <P and w along the characteristics are given by
Q ± <P =a constant (9.16)

where Q =-! J: 1
J(3- 4cos 2 w) 112 jsinw }dw

The equations (9.13) will have two distinct families of real characteristics,
that is, they will be hyperbolic if 3- 4cos 2 w > 0 and hence n/6 < w < Sn/6.
This range of points for which equations (9.13) are hyperbolic is shown
in heavy outline on the von Mises ellipse in figure 9.1(b).
Equations (9.13) have only one family of real characteristics, that is, will
be parabolic if 3- 4cos 2 w = 0 when the function, w, has the value of either
n/6 or Sn/6. If3- 4cos 2 w < 0 there are no real characteristics and equations
(9.13) are elliptic. The range of points on the von Mises ellipse in figure 9.1(b)
PLANE PLASTIC STRESS AND PSEUDO PLANE STRESS 315
corresponding to this type is shown in light outline. In the region of hyper-
bolicity it may be seen that um < -rmax' at the parabolic points um = -rmax'
and in the region of ellipticity um > -rmax.
For the hyperbolic region where 3 - 4 cos 2 w > 0 the function Q is given by
Q = - (n/4) +sin - 1 (2cos w/3 1' 2 )-(1/4)tan - 1 { (4cos w+ 3)/(3 -4cos 2 w) 1 ' 2 }
-(1/4)tan- 1 {(4cosw-3)/(3-4cos 2 w) 1' 2 } < 0 (9.17)
If another unknown function t/J(x, y) is introduced where
t/1 = - (n/2) + (1/2){ cos- 1 (cot w/3 1' 2 )} (9.18)
then the equations of the characteristics have the form
dyjdx = tan(c/> + t/1)
and Q + c/> =constant, C 1 , along an a characteristic
Also dyjdx = tan(c/>- t/1)
and n- c/> = constant, c2' along a f3 characteristic (9.19)
The characteristics of the differential equations of stress equilibrium
in the plastically deforming region are two families of curves which intersect
at an angle 21/1 as shown in figure 9.2 but are not orthogonal. The principal
directions bisect the angles between the characteristics such that the two
families of curves are inclined at angles of ± { (n/4) + (A./2)} to the direction
ofthe algebraic maximum principal stress u 1 where
(9.20)

a
Figure 9.2. Non-orthogonal a and Pcurved characteristics for a plane stress field
316 ENGINEERING PLASTICITY

it being assumed that a 1 > a 2 and a 3 = 0.


From the von Mises yield criterion (9.5) the principal stresses may be
determined as
a 1 = k(1+ 3 sin A.)/(1 + 3 sin 2 A.) 112 (9.21)
and a 2 = k(- 1 + 3 sin A.)/(1 + 3 sin 2 A.) 112
Equations (9.8) can be investigated to discover whether or not they
possess characteristics. Let the velocity components, u and v, and the compo-
nents of stress be given along a curve, C, where the cartesian (x, y) axes
are arranged to coincide, respectively, with the normal and tangent to the
curve C at any point P. If the velocity is continuous across the curve C,
oufoy and ovfoy are known at the point P. The partial differentials oufox
and ovfox at the point P can then be determined from equations (9.8) unless
2aY- ax= 0, in which case at the point P
(9.22)
The curve C will be a characteristic for the velocity equations if it coincides
at every point with a direction of zero rate of extension or contraction.
There are two such directions through any point. However, since there
will be a component of strain in the z direction they are generally not ortho-
gonal. The characteristics are inclined at angles of ± { (n/4) + (A./2)} to
the direction of the algebraic maximum principal stress, a 1, where
8y = {(81 + 82)/2} - {(81 - 82)/2} sin A.= 0
However 8tf82 = (2a1 - a2)/(2a2 - a1)
Hence sin A. = (8 1 + 82)/(8 1 - 82)
= (a 1 + a 2)/3(a 1 - a 2) (9.23)
The angle, A., is therefore real and the equations are hyperbolic when
am< k. It follows that when the von Mises yield criterion is used the charac-
teristics for the velocities coincide with those for the stresses.

9.3 SIMPLE STRESS STATES


The simplest solution is that for a uniform stress state. In a region of uniform
stress the field is generated by two nonorthogonal families of parallel
straight lines as shown in figure 9.3(a) along which there is no change in
the angle 1/J.
The particular type of field that applies to a given problem is determined
by the stress boundary conditions and the necessary velocity requirements.
For example, at a stress free boundary such as AB shown in figure 9.3(b)
the stress state is one of uniaxial tension or compression of magnitude
31' 2k acting parallel to the boundary. If the x axis is parallel to the boundary,
then ax= ± 3112 k, aY = •xy = 0 and the normal to the boundary coincides
PLANE PLASTIC STRESS AND PSEUDO PLANE STRESS 317

(a)

Strns free surface

(b)

Figure 9.3. (a) Plane stress slip line field for a uniform stress state; (b) plane stress
slip line field at a stress free surface and around a point of stress singularity

with one of the principal directions cp = 0. Assuming ux = + 31' 2 k (tensile)


then at the stress free boundary u x = u 1 , u Y =· u 2 = 0 and by referring to
figure 9.1(b) it will be seen that the value of w = n/3. Substituting this value
of win equation (9.18) gives t/1 = - (n/2) + 35° 16' = -54° 44', designating
the direction of an ex characteristic. Alternatively, from equation (9.20)
sin A= (u 1 + u 2 )/3(u 1 - u 2 ) = 1/3 where u 2 =0
or A= 19° 28'
. Therefore t/1 = ± { (n/4) + 9° 44'} = ± 54° 44'
to the direction of the algebraic maximum principal stress u 1 . The ex and
f3 characteristics are then designated as shown in figure 9.3(b). If, however,
the surface AB were not stress free, the angles at A and B would not necessarily
be equal but would depend on the nature of the loading on the surface AB.
Let AC in figure 9.3(b) be a boundary characteristic of the uniform stress
region ABC. If this region adjoins a region ACD having a different solution
318 ENGINEERING PLASTICITY

then all the characteristics of the family to which AC belongs will also be
straight lines. At the point of stress singularity, A, it is convenient to introduce
an auxiliary system of cylindrical coordinates r, e with pole at the point A
and polar axis AE.
The differential equations of equilibrium in cylindrical coordinates are
(oar/or)+ (ljr)(orr8 joe) + (ar- a8 )/r = 0 } (9.24)
(orr 8 /or) + (1/r)(oa 8 joe) + (2rr8 /r) = 0
and the von Mises yield criterion in cylindrical coordinates for plane stress is
(9.25)
The stress state that satisfies both the differential equations of equilibrium
(9.24) and the yield criterion (9.25) is given by
(Jr = kcos e, (J8 = 2kcos e, rr8 = ksin e (9.26)
where e is measured from the suitably chosen polar axis so that the stress
state is continuous across the boundary characteristic AC which separates
the two plastically deforming regions. The radial stress ar = kcos e is thus
constant along a radial characteristic. By considering the equilibrium of
the triangular element on the boundary characteristic AC it is found that
sine= (2/3) 112 (9.27)
or e = 54°44'
The polar axis AE must therefore be inclined at an angle e= 54° 4' to the
boundary characteristic AC or at an angle 2e to the stress free surface AB.
In the paper by Hill 5 , which is concerned with discontinuities in velocity
and stress, it is shown that the equation to curved characteristics of the
second family such as CD is given by
r 2 sin e= a constant (9.28)
Curved characteristics become asymptotic to the polar axis AE for which
e= 0. This is illustrated in figure 9.3(b) by the dash lines. Along AE the two
families of 11. and f3 characteristics converge into one and ar = k, a8 = 2k
which is the parabolic point w = n/6 on the von Mises ellipse in figure 9.1(b).

9.4 EQUILIBRIUM OF A THIN PLATE WITH A CIRCULAR


HOLE SUBJECTED TO UNIFORM PRESSURE
Consider a circular hole of diameter 2a in a thin infinite plate of an elastic-
perfectly plastic material and having a uniform thickness, h, to be subjected
to a gradually increasing pressure, p, at the edge of the hole where the radius
r = a, as shown in figure 9.4.
When the pressure, p, is low the plate is elastically deformed. The radial
and circumferential stresses are then given by
PLANE PLASTIC STRESS AND PSEUDO PLANE STRESS 319

Elastic

Rigid plastic annulus

Figure 9.4. The plastically deforming region surrounding an expanded hole in a


thin uniform plate showing thickening at the edge of the hole, the rigid-plastic annulus
and the radial and circumferential stress distributions

(9.29)
Consequently, when the deformation is elastic the stress state is one of
pure shear. Plastic yielding of the plate first occurs at the edge of the hole
when the pressure, p, attains the value of the yield shear stress, k, where
k = Y /3 1 12 according to the von Mises yield criterion. With increasing
pressure the plastically deforming region spreads through an annulus,
a ~ r ~ c, to some radius c. The stresses in the elastic region surrounding
this plastically deforming annulus where r ~ c are
ar = - k(cjr) 2 , a8 = + k(cjrf (9.30)
It is necessary to consider whether or not the plate begins to thicken
with the application of the pressure. The plate does not thicken near the
plastic boundary if the velocity equations are hyperbolic. This condition
is satisfied at the plastic boundary where p = - (ar + a8 )/2 = 0. The part
of the plastic region where the thickness of the plate is unaltered must
therefore extend over a finite annulus the inner boundary being a circle
where the velocity characteristics are coincident.
The radial and circumferential stresses are principal stresses and may
be written as
320 ENGINEERING PLASTICITY

u 6 = 2kcos{w- (n/6)}, ur = 2kcos{w + (n/6}} (9.31)


where u6 > ur.
The equation of equilibrium in the radial direction is
(do)dr) + (ur- u6 )/r = 0 (9.32)
By combining equations (9.31) and (9.32) the following differential equation
is obtained
(3 1 12 + cotw)dw + 2(drjr) = 0 (9.33)
Integration of equation (9.33) yields
r2 = exp(- 31i 2 w)(A/sin w) (9.34)
where A is an arbitrary constant.
At the boundary of the plastically deforming region, where r = c, the
stresses are continuous and the stress state is one of pure shear. Hence
when r = c, w = n/2 and the constant A= c2 exp(3 112 n/2). Therefore
(c/r) 2 = exp[3 1i 2 { w- (n/2)}] sin w (9.35)
The pressure, p, at the edge of the hole that produces the plastic region
of radius, c, is given by equation (9.31) for ur where r =a and w = W 8 • Hence
p = 2k cos {wa + (n/6)}
and (cjaf = exp[3 112 {wa- (n/2)}] sinwa (9.36)
It can be seen that W 8 ~ n/2 and increases with the ratio (c/a). The pressure,
p, attains a maximum value of 2k when W 8 = 5nj6. At the maximum pressure;
p = 2k, the characteristics coincide and the maximum radius, c, of the plastic
region is
cja = p = [ { exp(n/3 112 ) }/2] 112 R; 1.75 (9.37)
At the maximum pressure, p = 2k, the characteristics of the IX and pfamilies
become tangential, that is, they have zero inclination to the direction of the
algebraic maximum principal stress u6 ( 1/1 = 0). The characteristics thus
envelop the edge of the hole on the radius r = a. The pressure required to
generate the plastic region to the radius r = pa is equal to 2k and the circum-
ferential stress u6 = -kat the edge of the hole. Further increase in pressure
and expansion of the plastic region are not possible. At this maximum
pressure, p = 2k, thickening of the plate at the edge of the hole is free to occur
since the material inside the region of radius cjp R; 0.57c is not constrained
to remain rigid by the material at r > 0.57c. The thickening occurs because
the plastically deforming material cannot sustain a stress greater than 2k.
It follows that if the force exerted at the edge of the hole is increased then
the plate must thicken in this region to ensure that the stress does not exceed
the maximum possible value of 2k. The radial and circumferential stress
distributions are illustrated in figure 9.4.
PLANE PLASTIC STRESS AND PSEUDO PLANE STRESS 321
If elastic strains at the edge of the hole are neglected then from the Levy-
von Mises stress-strain rate relation (9.8) expressed in cylindrical coordinates
t 8 j(2a 8 - a,)= t,j(2a,- a8 ) (9.38)
Therefore { B8 ( a8 - 2a,)j(2a8 - a,)}+ t, = 0
Assuming the plate material to be incompressible
f.,+ f-8 + f.z = 0 (9.39)
The rate of relative thickening of the plate is then given by
tz = - ( t, + t8 ) = - t 8 ( a,+ a8 )/(2a8 - a,) (9.40)
At the maximum pressure, p = 2k, the stresses at the edge of the hole r = a
are a,= - 2k and a8 = - k. Substitution of these values into equation (9.40)
yields tz --+ oo, thus indicating that the thickness of the plate at the edge of
the hole is expected to increase freely for small deformation.

9.5 PLASTIC BENDING OF A THIN CIRCULAR PLATE


The problem of plastic bending of plates having cylindrical form and
subjected to lateral pressure is of increasing importance in manufacturing
industry. It can be considered as a case of plastic strain with axial symmetry
and, therefore, may have been considered in chapter 8. However, it is preferred
to include the topic in the present chapter because the treatment given is
that appropriate to a pseudo plane stress state.

9.5.1 Bending a thin circular plate using the von Mises yield criterion
Consider a thin circular plate of uniform thickness, 2h, and diameter, 2a,
to be supported at its outer edge in some manner and subjected to an axisym-
metric lateral pressure, p = p(r), where r is the radius as shown in figure
9.5(a). It is assumed that the plate material is rigid-perfectly plastic such
that the plate remains undeformed until the limiting value of the lateral
pressure is attained. As for the case of elastic bending of a plate, it is also
assumed that during plastic bending the median plane of the plate is neither
extended nor contracted and that planes initially normal to the median
plane remain plane.
In order that the usual convention for positive bending moments and
deflections is established, the positive z direction of the cylindrical coordinate
system (r, e, z) will be considered to be downwards. If the stress components
az, r,z be neglected by comparison with the stress components a,, a8 and
r,8 = r8 z = 0 then the plate is subjected to a plane stress state. For the element
of the plate shown in figure 9.5(b) the bending moment per unit length of
radius, M,, at radius, r, is given by

M, = f +h

-h
a,zdz (9.41)
322 ENGINEERING PLASTICITY
p

(b)

(c)

Figure 9.5. Bending of a thin circular plate showing the bending moments and
shear forces acting on an element at any radius r

At the section () = constant, the bending moment per unit length of


circumference, M 6 , is given by

(9.42)

In the circumferential direction the shear force must be zero for an axisym-
metric lateral pressure, that is, -rr6 = -c6z = 0, but in the radial direction the
shear force per unit length, F, at radius, r, is

(9.43)

which balances the force due to the external lateral pressure at this radius.
Therefore

2nrF + 2n f~ prdr = 0
PLANE PLASTIC STRESS AND PSEUDO PLANE STRESS 323

or F = - (1/r) J>rdr (9.44)

Considering the moments acting on the element of the plate shown in


figure 9.5(b) then for equilibrium
[ M. + {(oM./or )Dr} J(r +Dr )D()- M.rD()- 2MeDr sin ( D()/2)
+ [ F + {(oF /or )Dr} J(r +Dr )D()( Dr/2) + FrD()( Dr/2) = 0 (9.45)
if the moment of the external lateral pressure over the length, r, is neglected
as being of second order.
After simplifying equation (9.45) and neglecting second-order quantities,
the bending moments, M. and Me, and the shear force, F, are shown to be
related by the differential equation of moment equilibrium
(9.46)
Let the rate of bending of the plate be denoted by w = w(r). The strain-rate
components can then be stated as
(9.47)
where k. and ke are the coefficients of the rate of curvature of the median
plane of the plate in the radial and circumferential directions, respectively,
and are given by
(9.48)
The ratio of the strain rates, e.fee, is constant along the normal to the plane
of the plate. Neglecting elastic strain components, the Levy-Mises stress-
~train rate relations are given by equations (5.37), namely eii = iu'ii' where
A. is a scalar non-negative constant of proportionality. Expressed in cylindrical
coordinates
e.= l(u.- urn) where urn is the hydrostatic stress
= i[u.- {(u. + ue)/3}] since uz R:: 0
Therefore e.= (1/3)i(2u.- Ue) (9.49)
Similarly ile = (1/3)i(2ue- u.)
The non-negative stress components, u. and ue, must satisfy the von Mises
yield criterion for plane stress
u;- u.ue + u~ = 3k 2 = Y 2 (9.50)
Equations (9.49) can therefore be written in the form
e.= (1/3 )l( of I au.), ee (1/3 )i( of /oue) (9.51)
where f is the yield function given by the left-hand side of equation (9.50),
thus indicating that the strain-rate vector is normal to the von Mises yield
surface of figure 9.6(a).
324 ENGINEERING PLASTICITY
2 2 2
M-M M +M=Mp
r r e e

(a) (b)

-Me
(c)

Figure 9.6. (a) Representation of the von Mises yield criterion in the (ar, a8 )
plane for plane stress showing the strain-rate vector; (b) representation of equation
(9.53): M;- MrM8 + M; = M; in the (MrM8 ) plane corresponding to the von Mises
yield criterion in the (a r, a 8 ) plane showing the rate of curvature vector; (c) representation
of the equation Mr- M 8 = MP in the (Mr,M8) plane corresponding to the Tresca yield
criterion in the (ar, a 8) plane showing the various plastic regimes and the rate of curvature
vector

Equations (9.49) show that a 8 is proportional to ar along the normal.


However, the von Mises yield criterion (9.50) shows that ar = ± ¢ 1 (e./e 0 ) Y
and a0 = ± ¢ 2 (t.fe 0 ) Y, where ¢ 1 and ¢ 2 are some functions of the strain-rate
ratio F,./88 • The radial and circumferential stresses are therefore constant
along the normal for positive values of z and have opposite sign for negative
values of z, that is, they are discontinuous on the median plane of the plate.
This situation is analogous to that for the bending of a beam. The stresses
are shown as opposite points on the von Mises ellipse of figure 9.6(a).
For values of a. and a0 when z > 0 the bending moments are given as
(9.52)
PLANE PLASTIC STRESS AND PSEUDO PLANE STRESS 325
Combining equations (9.50) and (9.52) yields
M; - MrMB + Mi = M; (9.53)
where MP = Yh 2 is the maximum value of the bending moment required to
cause plastic deformation of the plate. Equation (9.53) is a special case of
the relation between forces and bending moments for plastic deformation
of shells as presented by Il'yushin 6 and, as could be expected, is represented
by an ellipse when plotted in the (Mr,M9) plane as shown in figure 9.6(b).
If the radial and circumferential stresses, ur, and u8 , are replaced by the
bending moments Mr and M 9 , respectively, in equations (9.49) and the strain-
rates by the coefficients of the rate of curvature kr and k9 , respectively, then
(9.54)
where 1/J is the expression on the left-hand side of equation (9.53) and A.* is
a scalar non-negative constant of proportionality. It follows that the vector
representing the rate of curvature is normal to the ellipse in the (M r, M 8 )
plane as shown in figure 9.6(b). Consequently, the Levy-von Mises associated
flow rule may be considered to be valid for the relations existing between the
bending moments M r and M 9 , and the coefficients of the rates of curvature
kr and k9 , as for the relations between the radial and circumferential stresses
ur and u8 and the strain-rates, er and 89 • Hence, by combining equations (9.53)
and (9.46) the bending moment, M 9 , may be eliminated and the following
non-linear differential equation for M. obtained
(dMr/dr) + (1/r)[(Mr/2) + {M;- (3/4)M;} 1 12 ] +F = 0 (9.55)
The solution of equation (9.55) with the appropriate boundary conditions
known enables the limiting value of the lateral pressure to be determined.
From equations (9.47) and (9.49) the differential equation for the rate of
bending ofthe plate is determined as
r(2M8 - Mr)(d 2 w/dr 2 ) - (2Mr- M 9 )(dw/dr) = 0 (9.56)
and this equation can be integrated if the bending moments Mr and M 9 are
known.
The boundary conditions vary depending upon the manner in which the
plate is supported. If the boundary is free then along the boundary M. = 0.
If the plate is simply supported, M r = 0 and w = 0, and if the plate is perfectly
clamped at its outer edge, w = 0 and dw/dr = 0, hence k8 = 0. Therefore
M.- 2M8 = 0 or Mr = 2M9 = ± (2/3 112 )MP.
The solution to equation (9.55) presents considerable difficulty. However,
for the case of a thin circular plate simply supported at its outer edge, the
numerical integration of equation (9.55) yields the limiting value of the
uniform lateral pressure, p*, as
(9.57)
326 ENGINEERING PLASTICITY

9.5.2 Bending a thin circular plate using the Tresca yield criterion

(a) Plate simply supported at its outer edge


The problem is simplified if the Tresca yield criterion is assumed to be
applicable, since the yield criterion is linear. The Tresca yield criterion is
represented by a hexagon in the (cr.,cr 9 ) plane or in the (M.,M 9 ) plane, as
shown in figure 9.6(c).
Consider a thin circular plate to be simply supported at its outer edge, as
shown in figure 9.7(a), and subjected to a uniform lateral pressure over a
circular area of radius c. The stresses, cr. and cr9 , attain their maximum values
at the centre of the plate where r = 0 and, therefore, at this point plastic
deformation first occurs. Thus, when r = 0, M. = M 9 = MP and one of the
plastic regimes BC, C, CD in figure 9.6(c) will be applicable. However, the
regime CD violates the differential equation of equilibrium (9.46) since for
CD, M r = M P and dM ./dr = 0. Also M 9 < M P, p > 0 and from the differential
equation of equilibrium it follows that dM.fdr < 0. For the same reasons, the
regime represented by the point C is not possible. The regime BC is possible
since M 9 = MP. From equation (9.46) with p = 0 when r > c
M = {(A/r)- (pr 2 /6)- MP when r ~ c (9.SS)
r (B/r)- (pc 2 /6)- MP when r ~ c

(b)

Figure 9. 7. (a) Thin circular plate simply supported at its outer edge subjected to a
uniform lateral pressure over a circular area of radius c; (b) thin circular plate perfectly
clamped at its outer edge subjected to a uniform lateral pressure over its entire surface
PLANE PLASTIC STRESS AND PSEUDO PLANE STRESS 327
where A and B are arbitrary constants. From the condition that M, is
bounded at the centre of the plate it follows that A = 0 and from the condition
that M, is continuous when r = c, B = pc 2 /3.
The bending moment, M,, decreases as the radius r increases, that is, the
point representing the state on the Tresca hexagon in figure 9.6(c) moves
from C to B. At the outer edge of the plate where it is simply supported,
M r = 0 and the regime B is realised, whilst for the remainder of the plate the
regime BC is applicable. Therefore, when r = a, M, = 0 and the limiting value
of the lateral pressure, p*, is given by
p* = [ 6a/{ c 2 (3b- 2c)} ]MP (9.59)
If the lateral pressure acts over the entire surface of the plate then c = a and
equation (9.59) reduces to
(9.60)
When the limiting lateral pressure, p*, is applied the Tresca associated flow
rule gives k, = 0, d 2 wfdr 2 = 0 for the regime BC. When r =a at the outer edge
of the plate
w = w 0 {1- (rfa)} (9.61)
where w0 is the value of the rate of bending at the centre of the plate which
remains unknown. However, this indicates that the mode of deformation is
such that the plate assumes the shape of the surface of a right circular cone
as illustrated by the dash line in figure 9.7(a).

(b) Plate perfectly clamped along its outer edge


Figure 9.7(b) shows the circular plate perfectly clamped at its outer edge
and, as in the previous case, the regime BC shown in figure 9.6(c) is applicable
near the centre of the plate. The bending moment, M,, decreases with
increasing radius and becomes zero at some radius r = p, say. The regime
BA becomes applicable up to the outer edge of the plate where r = a and
M, = - MP and this state is represented on the Tresca hexagon of figure
9.6(c) by the point A. When r ~ p equations (9.58) give
M, = MP- (pr 2 /6) (9.62)
When r = p,M, = Oand
(9.63)
When r > p, M 9 - M, = MP and from the equation of equilibrium (9.46)
Mr = MP ln(r/p)- {p(r 2 - p 2 )/4} (9.64)
since [MrJr=p = 0. When r =a, M, = - MP. Hence
5 + 2ln (a/p) = 3a 2fp 2 (9.65)
from which it can be calculated that p ~ 0.73a and the limiting lateral
328 ENGINEERING PLASTICITY

pressure, p*, is given by


p* ~ 11.3MP/a2 (9.66)
In the central region ofthe plate where r ::s:; p
d 2 w/dr 2 = 0, w = w0 + Cr
where w0 and C are constants. When r > p and the regime AB is applicable,
the associated flow rule gives kr: k0 = - 1: 1. Therefore (d 2 w/dr 2 ) + {(dw/
dr)/r} = 0. At the outer edge of the plate where r =a, w = 0 so that w = D
ln(r/a), where Dis a constant. This indicates that the condition dw/dr = 0
at the outer edge of the plate is violated and consequently a plastic hinge
is formed (see section 10.10).
The arbitrary constants, C and D, can be determined from the conditions
that w and dw/dr are continuous for r = p. As for the previous case, the
rate of bending at the centre of the plate, w0 , remains unknown but the
general nature of the mode of deformation of the plate is illustrated by the
dash line in figure 9. 7(b).

9.6 DEEP DRAWING OF A CIRCULAR SHEET METAL BLANK


The manufacturing process of forming a sheet metal blank into a cylindrical
cup- or box-shaped product known as deep drawing is characterised by the
use of a punch and die, the sheet being drawn radially inwards over the die
profile by the advancing punch. Deep drawing is an unsteady state process
in which the stress and strain distributions are very complex.
Consideration of the process will be restricted to only the simplest deep
drawing operation which is the manufacture of a cylindrical cup from a flat
circular blank of sheet metal and will be regarded here as a problem of
pseudo plane stress deformation. However, even for this case, no rigorous
analysis exists. The most extensive theoretical and experimental investiga-
tions concerned with the deep drawing of cylindrical cups were undertaken
in 1951 by Chung and Swift 7 • The analytical treatment is principally concern-
ed with radial drawing, plastic bending under tension and die profile frictional
resistance. The authors assume plane stress deformation of the flange and
also assume the principle of superposition to apply such that the separate
effects are considered to be additive. Although this may not be strictly
accurate it was shown that reasonably accurate predictions can be obtained.
A simple account of Swift's work on deep drawing has been given by
Hessenberg 8 and a useful summary in which the practical aspects are
emphasised was presented by Willis 9 •

9.6.1 Progressive stages in deep drawing a cylindrical cup


A conventional punch-die configuration for the manufacture of a cylindrical
cup from a flat circular sheet metal blank is shown diagrammatically in
PLANE PLASTIC STRESS AND PSEUDO PLANE STRESS 329

X y z y X

(a)

(b)

Figure 9.8. First stage deep drawing of a cylindrical cup: (a) zones of the blank;
(b) a partially drawn cup

figure 9.8. This includes a cylindrical flat-headed punch with a radiused


profile, a circular orifice-type die having a radiused profile and some form
of blank holder device. The lateral force provided by the blank holder may be
achieved by hydraulic or pneumatic pressure, by spring loading or by bolting
a mechanical blank holder to a predetermined fixed clearance. Blank holders
may therefore be designated as being either the constant pressure or constant
clearance type.
If a constant clearance exists between the blank holder and the die, equal
to the thickness of the blank material, then the flange of the drawn blank
deforms under essentially plane strain conditions. However, this is not likely
to be realised in practice because of the elastic deformation of the tooling
which occurs. On the other hand, if the blank holder force remains constant
and the mean lateral pressure exerted on the blank is small compared with
the current yield stress of the blank material then the flange is drawn under
conditions which approximate to plane stress.
The progressive stages in the deep drawing of a cylindrical cup are describ-
ed here following the approach used by Hessenberg. In figure 9.8(a) the
circular sheet metal blank is shown divided into three zones, X, Y and Z.
The outer annular zone, X, consists of material in contact with the die and
blank holder, the inner annular zone, Y, which is not initially in contact
330 ENGINEERING PLASTICITY

with either the punch or die and the circular zone, Z, which is in contact
with the head of the punch.
As the operation proceeds in the manner indicated in figure 9.8(b), which
shows a partially drawn cup, the material in zone X is drawn progressively
inwards towards the die profile under the influence of an applied radial
tensile stress and the effect of continually decreasing the radii in this zone
is to induce a compressive circumferential stress which causes an increase in
the material thickness. Unless a lateral pressure is applied by some form of
blank holding device, this compressive circumferential stress will cause the
flange to fold or wrinkle. Wrinkling is a condition of instability in compres-
sion, that is, buckling in the plane of the flange. The onset of wrinkling is
therefore likely to occur at an early stage in the radial drawing of the flange.
As the material in zone X passes over the die profile it is thinned by plastic
bending under tensile stress. Plastic unbending also occurs under tension as
the curved sheet is pulled straight to form the vertical cylindrical wall of
the cup between the punch and die throat. Finally, the inner part of zone X
is further thinned by tension between the die and the punch. The resultant
effect for the outer part of zone X is an increase in thickness ofthe material.
Considering the zone Y, then part is subjected to bending and sliding
over the die profile, part to stretching in tension between the die and the
punch head and part to bending and sliding over the punch profile.
Zone Z is subjected to stretching and sliding over the punch head, the
amount of strain depending on the geometry of the punch head and the
frictional conditions prevailing at the interface.
The various processes involved during the progressive stages of the deep
drawing of a flat-bottomed cylindrical cup may now be summarised as
follows:

(1) Radial drawing of the flange between the die and the blank holder
with frictional resistance present at the interfaces.
(2) Bending and sliding over the die profile under tension with frictional
resistance present at the interface.
(3) Unbending to form the vertical cylindrical wall ofthe cup.
(4) Stretching between the die and punch.
(5) Bending and sliding over the punch profile.
(6) Stretching and sliding over the punch head, the frictional resistance
depending on the prevailing conditions at the interface.

If the radial clearance between the punch and die throat is less than the
full thickness of the cup wall to be accommodated, then some degree of
ironing may also occur. Ironing is frequently employed to reduce the variation
in thickness of the cup wall. The clearance is reduced such that ironing of
the upper part of the wall is effected where the greatest amount of thickening
occurs.
The thickness history of an element of the blank originally in the flange
PLANE PLASTIC STRESS AND PSEUDO PLANE STRESS 331

- i:t1V'I/m2 A
I
~
c

r
r

(a) II IR
...
.."' -}~~r--.
...
"'I -r-
c:
u
·-.c::
I /
tc tB tB tA tA t t0
1-

Radius r -

Final cup
--thick nus
(b) X

--'Neck'N 1

---'Ntck' N 2

Figure 9.9. (a) Thickness history of an element during the deep drawing of a
cylindrical cup; (b) section through a deep drawn cylindrical cup using a flat headed
punch showing the formation of two 'necks' (thickness changes are exaggerated)

at a radius, R, where the thickness is, say, t 0 , is shown diagrammatically


in figure 9.9(a). When the current radius of the element is r, the thickness
of the metal is increased to t, due to radial drawing. At the point A, corres-
ponding to the commencement of the radiused die profile, thinning occurs
due to bending under tension as the material slides over the die profile. The
thickness of the material decreases from tA, to t~, say. This is followed by
further thickening to tB, as the material is further radially drawn during
the bending and sliding over the die profile from A to B. However, further
thinning to t~ occurs because of the unbending under tension to form the
vertical cylindrical cup wall. Additional thinning to some final value, tc,
will occur if the material in the region between B and C is subjected to
plastic stretch forming. Ironing will also cause further thinning should this
be employed.
332 ENGINEERING PLASTICITY
In zone Y of figure 9.8 there is a part which is not subjected to bending
and sliding over the die profile or the punch profile but is subjected to
tension. This particular material stretches and thins in tension to a lesser
extent than the material on either side. A thicker region therefore exists
between two apparent 'necks', N 1 and N 2 , as shown in figure 9.9(b), in
which the thickness changes are exaggerated. Instability in tension is likely
to occur at one or other of these points leading to the initiation of fracture
of the metal which is then propagated around the wall of the cup. Fracture
usually occurs at the point nearest to the punch head. The strains in the
region which is subjected to plastic stretch forming during deep drawing
with a hemispherical punch have been analysed by Woo 10 •

9.6.2 Radial drawing of the flange


Figure 9.10(a) shows a plan view of the circular sheet metal blank of initial
diameter, 2R 0 , and initial thickness, t 0 • The stresses acting on an element
of the blank during the radial drawing of the flange at any current radius, r,
are shown in figure 9.10(b). The radial and circumferential stresses are

(a)

Plan VJI:W

(b)

Figure 9.10. (a) Plan view of the circular sheet metal blank; (b) stresses acting on
an element of the blank at any current radius, r, during radial drawing of the flange
PLANE PLASTIC STRESS AND PSEUDO PLANE STRESS 333
ar and a9 , respectively, pis the lateral pressure exerted by the blank holder
and the frictional stresses at the blank holder-blank and blank -die inter-
faces are -r = Jl.p, where Jl. is the coefficient of friction. The physical conditions
are such that ar will be tensile whilst p and a9 are compressive. Also, since
the blank thickens most at the rim, it will be assumed that all the blank
holder force is concentrated at the rim and that elsewhere, p is zero. If a
metal blank holder is employed there is little thickening which is then mainly
due to elastic deformation of the tooling. The blank holder thus tends to
make the stress compressive only at the rim of the blank. In this case, the
radial drawing process can be regarded as one of plane stress deformation.
Resolving forces exerted on the element in a radial direction then for
equilibrium in this direction
(ar + dar)(r + dr)(t + dt)dO- arrtdO- 2a9 tdrsin(d0/2) + 2JJ.prd0dr = 0
(9.67)
After simplifying, dividing by rdOdr and neglecting small quantities
of higher order, equation (9.67) reduces to
{d(a.t)/dr} + {t(ar- a 8)/r} + 2JJ.p = 0
Since dtfdr is usually of the order of 2 x 10- 3 it will be assumed that
the thickening of the flange can be neglected so that
(dar/dr) + {(ar- a 8)/r} + 2JJ.p/t = 0 (9.68)
Hence, for pure radial drawing when frictional resistance at the blank
holder-blank and blank-die interfaces is absent
(9.69)
The radial and circumferential stresses must satisfy a yield criterion.
It is proposed to adopt the modified Tresca yield criterion, which was
previously discussed in section 8.8.2, as being a suitable compromise between
the Tresca and von Mises yield criteria. Hence
(9.70)
for a strain-hardening material, where Y is the mean yield stress of the
blank material and the coefficient m ~ 1.08.
Combining equations (9.69) and (9.70) produces r(dar/dr) +mY= 0
or dar= - mY(dr/r) (9.71)
Integration of equation (9.71) gives the radial stress required at any
radius, r, to produce frictionless pure radial drawing of the flange from
some initial radius, R. Therefore

ar = -mY f~ (drjr) = mYln(R/r) (9.72)


334 ENGINEERING PLASTICITY

9.6.3 Frictional resistance at the flange


IfF is the total blank holder force, then the radial stress required to overcome
the frictional resistance at the blank holder-flange and die-flange interfaces
is given by
(9.73)
It follows that the radial stress at the point A in figure 9.11 which is required
to radial draw the flange and overcome the frictional resistance at the blank
holder-flange and die-flange interfaces is
aA= ar(A) + af(A) = mYl~(R 0 /rA) + J1F/nR 0 t 0 (9.74)
Let the equivalent stress-strain (a, B) relation for the blank material
be given by
a= Aen (9. 75)
where eRJ e8 = ln (R/r), R, is the initial radius and r the current radius to the
annular element considered then the mean yield stress is

¥= J: J:
ade/ de

=A I: endeje
= Aen+ 1 /(n + 1)e
or Y = Ae /(n + 1)
0
(9.76)

Figure 9.11. Conventional punch-die configuration for the deep drawing of a


cylindrical cup showing the salient dimensions
PLANE PLASTIC STRESS AND PSEUDO PLANE STRESS 335
9.6.4 Plastic bending of the sheet, in tension, over the die profile
When the elements of metal in the flat flange reach the point A in figure 9.11
corresponding to the commencement of the die profile, they are formed
into a quarter section of a hollow toroid by bending under tension. The
metal is subsequently unbent at point B into the straight portion of the
vertical cup wall after passing over the die profile. Provided that bending
and unbending are each instantaneous, the metal elements have no freedom
to strain in the circumferential direction and, therefore, the corresponding
circumferential strain e9 = 0.
As a first approximation it will be assumed that the blank is subjected
to uniplanar bending under tension. The distributions of stress and strain
across the thickness of the blank when it is subjected to bending under
tension due to the radial drawing stress, u A' at A where the die profile
commences are illustrated in figure 9.12(a). For the conditions considered

Strain distribution

)
Applied
bending
moment

(a)

N N'
c c
N N

Final strain
~ Strain during bending
~ Additional strain during unbending

(b)

Figure 9.12. (a) Stress and strain distributions during bending under tension over
the die profile showing displacement of the neutral surface; (b) strain distribution
during unbending under tension
336 ENGINEERING PLASTICITY

it can be assumed that plane sections remain plane after bending. The
distribution of strain across the thickness of the blank is therefore linear,
the neutral surface of bending, NN, being displaced from the central surface,
CC, by an amount, A., which can be calculated by equating the resultant
force exerted on any transverse section to zero.
Assuming the material is strain-hardening having a mean yield stress, Y,
and the drawing stress at commencement of the die profile to be uA, so that
uAI Y = x, where 0 < x < 1, then for unit width of the sheet
- f(t/2)+.1. - - f(t/2)-.1.
Y dy+xYt- Y dy=O (9.77)
0 0

Therefore Y{(t/2) +A.}+ xYt- Y{(t/2)- A.}= o


or (t/2) +A.+ xt- (t/2) +A.= 0
and A.=- xt/2 (9.78)
For constancy of volume der + de11 + dez = 0. Therefore
dez = - der (9.79)
since de11 = 0. The equivalent incremental strain, de, is
de= [ (2/9) {(der- de11) 2 + (de11 - dez) 2 + (dez - der) 2 }] 112 (9.80)
By combining equations (9.79) and (9.80) it is then found that
de= (2/3 1' 2 )der (9.81)
The work done per unit width of sheet during bending of the element
which is distance y from the neutral surface NN
dW = stress x equivalent incremental strain x volume of element
= Y X de X (p + y)dOdy
= Y x (2/3 1i 2)der X (p + y)d0dy
but der ~ yf(p + y) (9.82)
where p is the radius of curvature of the neutral surface NN. Therefore
d w = Y(2/3 1' 2 )yd0dy (9.83)
The mean work done per unit width of sheet during bending over the
die profile whilst under tension is

w = (2/3 1' 2 )¥ do{J:' 2 >+.1. ydy + J:' 2 >-\dy}


where A.= - xt/2
= (2/3 1' 2 )Yd0[ {(t/2) + A.} 2 + {(t/2)- A.} 2 /2]
= (2/3 1' 2 ) Y d0t 2 { (1 + x 2 )/4} (9.84)
PLANE PLASTIC STRESS AND PSEUDO PLANE STRESS 337
Let the component of the drawing stress which is required to produce
the bending under tension be a b, then the work done per unit width of sheet
by this stress operating through the mean distance p c dO = {(t /2) + a} dO is
given by
(9.85)
where Pc is the radius of curvature of the central surface CC.
By combining equation (9.84) and (9.85) the additional stress required
to produce the bending under tension is derived as
ab::::; Yt(l + x 2)/(2.3 1i 2){ (t/2) +a}
However, a= rd, the radius of the die profile x = lT A/Y and lTA= lTr<A> + lTr<A>·
Therefore
lTb::::; Yt 0 {1 +(aA/Y) 2 }/(2.Jl12 ){rd +(t0/2)} (9.86)
if the change in thickness of the material during bending is neglected.

9.6.5 Additional radial drawing as the material passes over the die profile
At the same time as the material is subjected to plastic bending under
tension when passing over the die profile from A to B, it is also subjected
to further radial drawing when the radius of the elements is reduced from
rA to rP, say, shown in figure 9.11 where rA= r1 + rd and 2r1 is the throat
diameter of the die, the effective radius of the punch rP = rP + (t 0 j2) and
2rP is the diameter of the punch.
The increase in stress required to produce this radial drawing as the
material passes over the die profile from A to B is then given by
(9.87)

9.6.6 Frictional resistance at die profile


The increase in stress required to overcome the frictional resistance at the
die profile, O"rcA-+Bl' as the material slides from A to B can be estimated by
considering the problem to be analogous to a belt passing over a pulley
with Coulomb friction present at the interface. If T8 is the tension at B
and TA is the tension at A then
(9.88)
where Jl is the coefficient of friction at the interface and 0 ::::; n/2. In reality,
0 is usually less than n/2.
For a small die profile radius, rd, the cross-sectional areas of the blank
at radii, rA and rP, are approximately equal. Therefore
{aA+ ab + O"r(A-+Bl} /(a A+ a b)::::; exp(JLn/2)
or af(A-+B) ::::; (a A+ a b){ exp(JLn/2)- 1} (9.89)
338 ENGINEERING PLASTICITY

9.6. 7 Unbending of the sheet at die prof'Ile


The exact analysis for unbending after the material has passed over the
die profile will be even more complex than that for bending because of the
different strain histories of the elements across the thickness of the material
due to the plastic bending operation.
The strain distributions of elements during the bending and unbending
operations under tension are shown diagrammatically in figure 9.12(b). The
neutral surface during bending is NN and N'N' is that for unbending. It
will be appreciated that the outer layers of the material are subjected to a
greater strain than the inner layers, to which are added the changes in
strain due to the additional radial drawing as the material passes over the
die profile from A to B. However, it is considered sufficiently accurate to
use equation (9.86) as a first approximation for the unbending stress, uu,
provided that uB, the drawing stress required before unbending, is substituted
for u A. Therefore
O'u ~ Yt 0 {1 + (uB/Y) 2 }/(2.3 112){rd + (t 0 /2)} (9.90)
where the drawing stress before unbending
(9.91)

9.6.8 Estimation of wall stress and punch force required to deep draw
a cylindrical cup
Assuming that the various effects are additive then the total wall stress in
the drawn cup at radius rP is
O'w ~ O'r(A) + O'f(A) + O'b + O'f(A-+B) + O'r(A-+B) + O'u (9.92)
and the punch force, P, is given by
(9.93)
This approximate analysis excludes any stretch forming which may occur
in the region between B and C during the forming of a cylindrical cup and
assumes that there is no relative movement between the material and the
punch head, that is, the punch head is perfectly rough. Engineering metals
which are suitable for deep drawing are usually anisotropic to some degree.
The analysis neglects the effects of anisotropy. Also no allowance has been
made for ironing should this occur. Nevertheless, the author has found that
a good correlation exists between the predicted and experimental values
of punch force required for the deep drawing of cylindrical cups from 70/30
brass and extra deep drawing quality mild steel sheet. As might be anticipated
it is necessary to achieve a high degree of accuracy in determining the mean
yield stress, Y, for the sheet metal. An elementary analysis was attempted
by Duncan and Johnson 11 , in 1969, to predict the punch force required
to deep draw a cylindrical cup. A useful survey of the theory of deep drawing
by Alexander 12 should also be referred to.
PLANE PLASTIC STRESS AND PSEUDO PLANE STRESS 339
9.6.9 Ironing of the cup wall
During the deep drawing of a cylindrical cup the wall tends to become
thicker than the original blank as a result of the radial drawing of the flange.
In a first-stage drawing process such as that considered, the upper two-thirds
of the cup wall is likely to be affected and the thickness may be increased
over part of the wall by as much as 30 per cent. It is therefore sometimes
desirable to subject the cup to ironing for the following main reasons:
(a) to produce a more uniform wall thickness; (b) to remove evidence of
incipient wrinkling which may have occurred; and (c) to assist in straightening
the cup wall which may have become bell-mouthed at the open end.
The circumstances in which ironing is used in practice were discussed by
Barlow 13 , and Hill 14 has analysed the ironing of a thin-walled cup as a steady
state approximate plane strain process by means of a slip line field analogous
to that for plane strain sheet drawing through a wedge-shaped die with
Coulomb friction present at the workpiece-die interface. A technological
theory of ironing was developed by Sachs, Lubahn and Tracy 15 for a non-
strain-hardening material. The theory was extended by Swift et al. to include
frictional resistance, strain-hardening of the workpiece material and redun-
dant deformation. Ironing of cups when the wall thickness is not small
compared with the cup diameter has been studied by Briggs and Swift 16 •
In the experiments by Swift and his associates, the ironing force-fractional
reduction characteristics for various die angles and lubrication conditions
were obtained. These investigations were extended in 1953 by Freeman
and Leeming 1 7 who measured the two components of the ironing force
separately and together. The two components are the punch head force and
the force required to overcome the frictional resistance along the punch
surface. The punch head force determines the tension in the cup wall and
consequently limits the maximum fractional reduction possible for a given
punch force. A more recent theoretical study of ironing has been presented
by Fukui and Hansson 18 . The relevant properties of materials in terms of
their suitability for deep drawing have been examined by Jevons 19 •
An estimate of the additional punch force required to effect ironing of
a cylindrical cup wall can be obtained by reference to a slip line field solution
in the manner proposed by Hill. It is assumed that the thickness of the cup
wall is to be reduced but the internal diameter remains unchanged and
equal to the diameter of the punch, d = 2rP. Because the cup wall is elongated
during ironing there is frictional resistance at the punch-cup interface in
addition to that at the cup-die interface. However, the effect of the frictional
resistance at the punch-cup interface is less significant than that at the
cup-die interface and is therefore neglected.
The circumferential strain of an element in the cup wall at the punch-cup
interface, that is, at radius r = r P, is zero but increases for elements at greater
initial radii. However, for a thin-walled cup when the diameter of the cup
is, for example, not less than 30 times the thickness of the cup wall to be
ironed, the circumferential strain approximates to zero. It is then sufficiently
340 ENGINEERING PLASTICITY
accurate to treat the process as one of plane strain deformation and the
problem is then identical with that of plane strain sheet drawing through a
wedge-shaped die of semi-angle, a, with Coulomb friction present at the
workpiece-die interface. The surface of the punch is regarded as the friction-
less central plane of the sheet. Figure 9.13(a) shows the ironing of a cylindrical
cup to reduce the wall thickness from h 1 to h2 , and in figure 9.13(b) is given
the appropriate slip line field when the deformation is treated as approxima-
tely plane strain for a moderate fractional reduction. The normal die pressure
with Coulomb friction present at the cup-die interface is q', and the frictional
stress is -r = Jl.q', having the sense shown where Jl. is the coefficient of friction.
The angle, l'f, is less than n/4 the value depending on the value of Jl. and cos 11 =
M'fk.
By resolving the forces exerted on an element in the plastically deforming
region of the cup wall in the axial direction and equating to zero for equili-
brium, the mean ironing stress, t', can be determined as follows:
t'h 2 - q' { (h 1 - h2 )/sin a} sin a - JJ.q' { (h 1 - h2 )/sin a} cos a = 0 (9.94)
or t' = q' { (h 1 - h2 )/h 2 } + JJ.q' { (h 1 - h2 )/h 2 } cot a

(a)

(b)

Figure 9.13. (a) Ironing of a cylindrical cup wall; (b) slip line field solution for
ironing of a cup wall assuming plane strain deformation and Coulomb friction present
at the cup-die interface
PLANE PLASTIC STRESS AND PSEUDO PLANE STRESS 341
However, the fractional reduction, R = (h 1 - h2 )/h 1 and (h 1 - h2 )/h 2 =
R/(1 - R). Therefore
t' = q' {R/(1- R) }(1 + J.LCOt IX) (9.95)
If t and q are the values of the mean ironing stress and normal die pressure,
respectively, when the cup-die interface is frictionless such that J.L = 0, then
t = q{R/(1- R)} (9.96)
The normal die pressure is not significantly affe9ted by the frictional
resistance for values of J.L < 0.1, in which case the approximation q' ~ q can
be made. Therefore
t' ~ (1 + J.L cot et)t (9.97)
The additional punch force required to effect the ironing of the cup wall
is then given by
F; = nh 2 (2rP + h2 )t' ~ nh 2 (2rP + h2 )(1 + J.LCOtet)t (9.98)
The frictionless ironing stress, t, is determined by resolving the forces
due to the hydrostatic pressure distribution and the shear stress along the
corresponding f3 exit slip line BCD when 11 = n/4 in the axial direction and
equating to the ironing force per unit width, th 2 , following the method
outlined for cutting in section 7.4.4.

9.6.10 Re-drawing of cylindrical cups


An indication of the extent to which a circular sheet metal blank can be
formed into a cylindrical cup-shaped product is the drawing ratio. The
drawing ratio is defmed as the ratio of the blank diameter to the throat
diameter of the die. In manufacturing cups from flat circular blanks of the
common engineering metals, the maximum first-stage drawing ratio is
usually about 2.0, corresponding to a cup height to diameter ratio of about
unity.
However, it was shown by Wallace 20 that the limiting drawing ratio
can be increased by increasing the frictional resistance between the blank
and the punch. By producing a knurled fmish on the punch surface, the
limiting drawing ratio was increased from 2.325, when using a polished and
well-lubricated punch, to 2.55 for fully stabilised, extra deep drawing quality
mild steel. Even higher drawing ratios have been achieved by Kasuga and
Tsutsumi 21 and also by El-Sebaie and Mellor 22 when deep drawing into a
high pressure medium, which had the effect of keeping the partly drawn cup
in intimate contact with the punch under high hydrostatic pressure. The
same principle is involved in the Schuler process 23 and the hydroform
process24 to increase the depth of draw.
The limiting drawing ratio was shown by Whiteley 25 to also depend on
the normal anisotropy in the sheet metal as characterised by the R value.
342 ENGINEERING PLASTICITY

When a material having a high R value is deep drawn, the change in thickness
of the flange as it is drawn radially inwards over the die profile is less than
that for a material having a low R value. The thickness ofthe cup wall is then
more uniform. For a given blank diameter a deeper cup will therefore result
when using a high R-value material. The effect ofthe strain-hardening index,
that is, the n value, and also the R value on the limiting drawing ratio has
been studied by El-Sebaie and Mellor 26 •
In conventional practice, if a cup is to have a height to diameter ratio
greater than unity such that the first-stage limiting drawing ratio is to be
exceeded, the process of re-drawing in one or more stages becomes necessary.
The methods of re-drawing may be classified into two main groups, namely,
(a) direct re-drawing and (b) reverse re-drawing which are illustrated diagram-
matically in figure 9.14. In the direct method, shown in figure 9.14(a), it will
be seen that the external surface of the cup produced in the first-stage drawing
process remains the external surface of the re-drawn cup. However, in the
reverse method, shown in figure 9.14(b), the cup is turned inside out such

(a)

(b)

Figure 9.14. Diagrammatic representation ofthe two main methods of re-drawing


cylindrical cups: (a) direct re-drawing; (b) reverse re-drawing
PLANE PLASTIC STRESS AND PSEUDO PLANE STRESS 343
that the internal surface of the first-stage cup becomes the external surface
of the re-drawn cup. The important difference between the direct and reverse
methods of re-drawing is the number of bending and unbending operations
involved. In reverse re-drawing one bending and unbending operation over
the die is avoided. It is interesting to note that metals which are very suitable
for first-stage drawing are not always as suitable for re-drawing. For first-
stage drawing, materials having a high rate of strain-hardening give good
results. On the other hand, materials which have a low rate of strain-harden-
ing are more suitable for re-drawing assuming that inter-stage annealing is
not employed.
Experimental re-drawing investigations were undertaken by Chung and
Swift 27 on brass, mild steel and aluminium first-stage flat-bottomed cylindri-
cal cups of 4 in diameter using an experimental crank press. A theoretical
analysis of the re-drawing of cylindrical cups has been presented by Fogg 28 •

9.6.11 Earing
The formation of ears at the rim of a deep drawn cylindrical cup as a mani-
festation of the effect of planar anisotropy was briefly discussed in section
6.5.3. Baring is undesirable since it represents a wastage of material, and the
elimination of ears to produce a plane rim involves a secondary machining
operation which increases manufacturing costs. However, anisotropy
may be usefully exploited in some asymmetric drawing operations 29 •30 •
During the deep drawing of a cylindrical cup, ears could be expected to
develop from the rim ofthe blank at positions where the uniaxial yield stress
has a minimum value in the circumferential direction. This is found to be the
case for certain steels and commercially pure aluminium when four ears are
developed. The behaviour of some brasses for which six ears may develop
is more complex. The macroscopic theory for the earing of deep drawn cups
advanced by Hill 31 can explain the formation of four ears. One of the more
successful attempts in predicting earing behaviour is due to Tucker 32 who
adopted a crystallographic approach to describe the earing of cups produced
from single crystals of aluminium. In a recent paper by Jimma 3 3, slip line
field theory was used to predict earing behaviour. More recently, Sowerby
and Johnson 34 have developed anisotropic slip line fields in the flange of a
deep drawn cylindrical cup to predict the location of ears at the onset of
radial drawing assuming the deformation to be plane strain.

9.6.12 Flange wrinkling during deep drawing


The outer part of the circular blank is subjected to a tensile radial stress as
the flange is formed and a compressive circumferential stress is induced.
When the magnitude of these stresses attain a critical value, depending on
the current flange dimensions, a condition of instability in compression,
that is, buckling occurs when the flange material collapses into waves or
344 ENGINEERING PLASTICITY

wrinkles in the circumferential direction. Since buckling is associated with a


transition phenomenon between the elastic and plastic states, wrinkling of
the flange usually first occurs early in the radial drawing process. Wrinkling
is controlled by changing the stress state of elements in the flange which is
achieved by subjecting the blank to a lateral pressure using some form of
blank holder.
The phenomenon was examined by Geckler 35 , in 1928, and an extensive
experimental and theoretical study of flange wrinkling was undertaken by
Senior 36 , in 1956. It was shown theoretically that, in the absence of a blank
holder and therefore when no lateral pressure is exerted on the flange, the
onset of wrinkling occurs in the range
0.46(t 0 /D 0 ) ~ a6 /E 0 ~ 0.58(t 0 /D 0 ) (9.99)
and the number of waves or wrinkles formed, n, is within the limits
1.65(a/b) ~ n ~ 2.08( afb) (9.100)
where a6 is the induced circumferential stress, t 0 , and D 0 = 2R 0 are the initial
thickness and diameter of the circular blank, respectively, a is the mean
flange radius, b is the flange breadth and E 0 is the plastic buckling modulus
given by
(9.101)

-A-
-~-2

-~-3

-~-4
Figure 9.15. Illustrating how the number of waves or wrinkles may be increased
by a reverse buckling process during radial drawing of the flange
PLANE PLASTIC STRESS AND PSEUDO PLANE STRESS 345
where E is the modulus of elasticity and P is the appropriate slope of the
true stress-natural strain characteristic curve for the blank material.
As previously stated, the wrinkling of flanges is controlled in deep drawing
practice by employing a blank holder to exert lateral pressure on the blank
and the development of wrinkles of large amplitude is then prevented by
moderate lateral pressure. However, the number of waves may be increased
by a reverse buckling process, as illustrated diagrammatically in figure 9.15.
The more numerous wrinkles of small amplitude which may develop are
easily ironed to produce an acceptable finish at the rim of the cup without
a high punch force being necessary.
The lateral pressure is effected in the constant clearance type of blank
holder by the use of compression springs or by the elastic deflection of the
holding-down bolts of the blank holder. The blank holder force then varies
linearly with the blank holder deflection and hence the amplitude of the
waves or wrinkles. The number of waves formed is independent of the wave
amplitude and is given by
n = a(3.8s/E 0 bat 3 ) 114 (9.102)
where s is the lateral stiffness of the blank holder and t the current mean
thickness of the flange. In the case of the constant pressure type of blank
holder the theory suggests that the number of waves is dependent on the
amplitude of the waves and is given by
n = [ {1.34E0 a 4 t 3 + (2.5a 3 b3 F/A.) }/E0 b4 t 3 Jl' 4 (9.103)
The amplitude of the waves is A. and F is the total blank holder force.
It was shown by Senior that, for various materials, there is a minimum
blank holder force required to control the amplitude of the wrinkles to an
acceptable level and that the punch force required to produce the cup is not
significantly affected. However, there is a critical diameter of blank above
which wrinkling occurs iflateral pressure is not exerted on the blank by some
form of blank holder mechanism.

REFERENCES
I. Hodge, P. G. Jnr., The method of characteristics applied to the problems
of steady motion in plane plastic stress, Q. appl. Math., 8, 381 (1950)
2. Hodge, P. G. Jnr., Yield conditions in plane plastic stress, J. Math.
Phys., 29, 38 (1951)
3. Hill, R., The Mathematical Theory of Plasticity, ch. XI, p. 300, Oxford
University Press, London (1950)
4. Prager, W., On the use of singular yield conditions and associated flow
rules, J. appl. Mech., 20, 317 (1953)
5. Hill, R., On discontinuous plastic states with special reference to
localised necking in thin sheets, J. Mech. Phys. Solids, 1, 19 (1952)
346 ENGINEERING PLASTICITY

6. Il'yushin, A. A., Plasticity, (in Russian) Gostekhizdat, USSR (1948)


7. Chung, S. Y. and Swift, H. W., Cup drawing from a flat blank-Part I,
Experimental Investigation, Part II, Analytical Investigation, Proc.
Instn mech. Engrs, 165, 199 (1951)
8. Hessenberg, W. C. F., A simple account of some of Professor Swift's
work on deep-drawing, B.I.S.R.A. Rep. MW/1954 (1954)
9. Willis, J., Deep Drawing: A review of the practical aspects of Professor
H. W. Swift's researches, Butterworths, London (1954)
10. Woo, D. M., On the complete solution of the deep-drawing problem,
Int. J. mech. Sci., 10, 83 (1968)
11. Duncan, J. L. and Johnson, W., Approximate analysis of loads in
axisymmetric deep drawing, in Proc. 9th Intern. Mach. Tool Des. and
Res. Conf., p. 303, Pergamon Press, Oxford (1969)
12. Alexander, J. M., An appraisal of the theory of deep drawing, Metal/.
Rev., 5, 349 (1960)
13. Barlow, D. A., The formability of aluminium alloys, Engineering,
Lond., 181, 329, 366 and 393 (1956)
14. Hill, R., The Mathematical Theory of Plasticity, ch. VII, p. 178, Oxford
University Press, London (1950)
15. Sachs, G., Lubahn, J.D. and Tracy, D.P., Drawing thin-walled tubing
with a moving mandrel through a single stationary die, J. appl. Mech, 11,
199 (1944)
16. Briggs, G. C. and Swift, H. W., Ironing of thick-walled cups, Motor
Ind. Res. Assoc. Rep. R/4 (1947)
17. Freeman, P. and Leeming, H., Ironing of thin-walled metal cups-the
distribution of punch load, B.I.S.R.A. Rep. No. MW/E/46/53 (1953)
18. Fukui, S. and Hansson, A., Analytical study of wall ironing considering
work-hardening, Annals C.I.R.P., 18, 593 (1970)
19. Jevons, J.D., The Metallurgy of Deep Drawing and Pressing, 2nd ed.,
Chapman and Hall, London (1949)
20. Wallace, J. F., Improvements in punches for cylindrical deep drawing,
Sh. Metal Inds, 37, 901 (1960)
21. Kasuga, Y. and Tsutsumi, S., Pressure lubricated deep drawing, Bull.
Japan Soc. mech. Engrs, 8, 120 (1965)
22. El-Sebaie, M.G. and Mellor, P. B., Plastic instability conditions when
deep drawing into a high pressure medium, Int. J. mech. Sci., 15, 485
(1973)
23. Buerk, E., Hydro-mechanical drawing, Sh. Metal Inds, 44, 182 (1967)
24. Lay, 0. P., Some recent developments in the Hydroforming process,
Sh. Metal Inds, p. 659 (September, 1971)
25. Whiteley, R. L., The importance of directionality in drawing quality
sheet steel, Trans. Am. Soc. mech. Engrs, 52, 154 (1960)
26. El-Sebaie, M.G. and Mellor, P. B., Plastic instability conditions in the
deep-drawing of a circular blank of sheet metal, Int. J. mech. Sci., 14,
535 (1972)
PLANE PLASTIC STRESS AND PSEUDO PLANE STRESS 347
27. Chung, S. Y. and Swift, H. W., An experimental investigation into the
re-drawing of cylindrical shells, Proc. Instn mech. Engrs, lB, 437 (1952)
28. Fogg, B., Theoretical analysis for the redrawing of cylindrical cups
through conical dies without pressure sleeves, J. mech. Engng Sci., 10,
141 (1968)
29. Lankford, W. T., Snyder, S. 0. and Bauscher, J. A., New criteria for
predicting the press performance of deep drawing sheets, Trans. Am.
Soc. mech. Engrs, 42, 1197 (1950)
30. Lloyd, D. H., Metallurgical engineering in the pressed metal industry,
Sh. Metal Inds, 39, 82 (1962)
31. Hill, R., The Mathematical Theory of Plasticity, ch. XII, p. 328, Oxford
University Press, London (1950)
32. Tucker, G. E. G., Texture and earing in deep drawing of aluminium,
Acta me tall., 9, 275 (1961)
33. Jimma, T., Earing of deep drawn cylindrical cups (in Japanese), Plastic
properties and machining, 11, 707 (1970)
34. Sowerby, R. and Johnson, W., Prediction of earing in cups drawn from
anisotropic sheet using slip-line field theory, J. Strain Analysis, 9 (2), 102
(1974)
35. Geckler, J. W., Plastic folding of the walls of hollow cylinders and some
other folding phenomena in bowls and sheets, Z. angew. Math. Mech.,
8, 341 (1928)
36. Senior, B. W., Flange wrinkling in deep-drawing operations, J. Mech.
Phys. Solids, 4, 235 (1956)
10
Extremum Principles for a
Rigid-Perfectly Plastic Material

10.1 GENERAL CONSIDERATIONS


In the theory of plasticity a number of general theorems have an important
role and the most important of these are the theorems on the extremum
properties of a solution or extremum principles and the so-called uniqueness
theorems. These theorems permit the possibility of the direct construction of
solutions which do not demand the integration of the differential equations
of equilibrium. This possibility is very important because in plasticity the
problems are usually nonlinear.
It has been previously indicated, for example in section 7.4.2 that difficulties
can arise in relation to the rigid-perfectly plastic model due to the non-
uniqueness of the solution when more than a single valid slip line field
solution for the problem exists. This situation compels the introduction of
kinematically admissible velocity fields and statically possible stress distri-
butions together with the formulation of a criterion for selection. The
extremum principles for a rigid-perfectly plastic material are therefore of
considerable significance. These lead to an effective method of determining
the limiting value of, for example, the force, pressure or couple to cause
unconstrained plastic deformation using successive approximations by
means of upper and lower estimates. The former is known as an upper bound
estimate which is expected to be an overestimate of the limiting value and
the other, referred to as a lower bound estimate, is certainly an underestimate.
Consequently, an upper bound estimate is of special importance for engineers,
particularly those concerned with metal forming processes, since it generally
provides an estimate for the required forming parameter which ensures that
the metal forming operation can be performed, unlike the lower bound
estimate which does not.
A lower bound estimate requires a statically admissible stress field to
be determined throughout the entire material and stress discontinuities are
permitted. However, there is no attempt to ensure that the velocity conditions
are satisfied at any point in the deforming material. The alternative approach
in obtaining an upper bound estimate is to establish a kinematically admis-
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 349
sible velocity field in which discontinuities in the tangential component of
velocity are permitted but the stress equilibrium equations may not be
satisfied in some regions. Should upper and lower bounds coincide then it can
be asserted that the value obtained is exact, and from the uniqueness theorem
it then follows that a complete solution has been determined if the statically
admissible stress field and the kinematically admissible velocity field are
related according to the plastic potential flow rule.
Application of energy theorems are presented in the book by Gvozdev 1
who apparently produced the earliest proof of the theorems of limit analysis.
A translation of his paper, published in 1936, is given by Haythomthwaite2 •
However, the theorems are also deducible from the principles advanced by
Hill 3 • A rigorous account of the general theorems of elastic-plastic media
as well as a detailed bibliography can be found in the work of Koiter4 •
Drucker, Greenberg and Prager 5 stated the limit theorems for an elastic-
perfectly plastic material from which the expressions for bounds can be
determined.
In this chapter, reasonably rigorous proofs are given of the extremum
principles and the limit theorems applied to a rigid-perfectly plastic material
where the presentation involves tensor notation. It is therefore suggested
that the reader revises sections 5.7 and 5.8 before studying this chapter.
For the proofs given, velocities, vi, and plastic strain-rates, tip are referred
to throughout, associated with the rate of energy dissipation or rate of
external work done. The theorems can also be presented in terms of displace-
ment increments, du, and plastic strain increments, de, and the increment of
work done, which was the case in section 5.8.

10.2 BASIC ENERGY EQUATION OR VIRTUAL WORK


EQUATION
The differential equations of force equilibrium were derived in section 2.2.1
and when stated in tensor notation are
(10.1)
where crii is the stress tensor and Xi are the components of the body force
per unit volume.
It is assumed here that inertia forces are absent or insignificant. Many
problems in plasticity theory may be regarded as quasi-static, that is, the
inertia forces due to plastic flow may be neglected. The problem may be
considered as quasi-static if the dimensionless parameter p V 2 /Y ~ 1, where p
is the density of the material, V is a characteristic velocity and Y the uniaxial
yield stress.
Consider a body which occupies a volume, V, and is bounded by the
surface, S, as shown in figure 10.1. The surface, S, is considered to comprise
two parts, SF and Sv. Assume crii to be any stress field which satisfies the
differential equations of equilibrium (10.1) which is consistent with the
350 ENGINEERING PLASTICITY

T n

Total surface area 5

Figure 10.1. A body of total volume, V, bounded by a surface of total area,


S, which is subjected to a stress field, uii, producing a traction, T, on the part surface,
SF, and also a prescribed velocity, v, on the part surface, Sv

prescribed tractions, y;, on the surface, SF, ofthe body and body forces X;.
Then
(10.2)
where ni are the direction cosines of the outward unit normal n.
On the other part, Sv, of the surface let the velocity, v, be prescribed
and its components be denoted by v;. Corresponding to this field the strain-
rate components are li;i = dei/dt, the differentiation following the particle,
and are given by

8.. = (1/2){_!_(du;) +_!__(~)}


IJ oxj dt OX; dt

(10.3)

where t denotes time and V; = du;/dt is the velocity of the particle with
current position vector X;. These stress and velocity fields in other respects
are arbitrary and, in general, are not interrelated.
It is assumed for the present that these fields are continuous, that is, no
stress or velocity discontinuities exist. However, these restrictions will be
removed later when surfaces of stress and velocity discontinuity are consi-
dered. The configuration of the body is assumed to be characteristic of
a yield state or else the configuration is negligibly different from its initial
state, then V and S are the volume and surface area up to the deformation.
For the complete body the rate of external work done can be stated as

(10.4)
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 351
where the first integration extends over the whole surface S and the other
integration over the whole volume of the body V. However, the surface
integral

Therefore

The first integral on the right-hand side of the above equation is equal to
zero which follows from the differential equation of equilibrium (10.1).
Hence
(10.5)

which is known as the virtual work equation.

10.3 SURFACES OF STRESS AND VELOCITY DISCONTINUITY


Equation (10.5) in the previous section assumes continuous stress and
velocity fields. However, discontinuities in stresses occur frequently in
plastic deformation and, for the rigid-perfectly plastic material, velocity
discontinuities are also inevitable. It is therefore necessary to consider the
generalisation of the virtual work equation.

10.3.1 Surfaces of stress discontinuity


Consider the case where the stresses are discontinuous at certain surfaces,
say Sk(k = 1, 2, 3 ... ). These surfaces divide the body into a fmite number
of parts in each of which the stress varies. The respective surface integrals
are performed over the surface for each individual region. Let the tractions
rP>
I
=aIJ.. n~l)
J
act on one side of the surface Sk and T!I 2 >=aIJ.. n!l>
J
act on the
other side.
The equilibrium requirement for an element at a surface of stress
discontinuity is
(10.6)
352 ENGINEERING PLASTICITY

Consequently, if all such equations for each region of the body are added
it is then found that all the integrals over the surface of stress discontinuity
cancel. This means that the normal stress components, uii ni, must be conti-
nuous across the surface, that is, the existence of stress discontinuities does
not affect the virtual work equation. Equation (10.5) is thus valid when the
equilibrium stress field, uii, contains surfaces of stress discontinuity.

10.3.2 Surfaces of velocity discontinuity


Assume the body under consideration to be divided into two regions, (1)
and (2), by a surface of velocity discontinuity, S0 , where the velocity field
vi is otherwise continuous, as shown in figure 10.2(a). It should be noted
that for metals, when incompressibility is assumed during plastic deforma-
tion, the component of velocity normal to the surface of velocity disconti-

z y

(b)

Figure 10.2. (a) A surface of velocity discontinuity, S0 , which separates region (1)
from region (2) in a body across which there is a tangential velocity discontinuity V*;
(b) the surface of velocity discontinuity, S0 , as the limiting position of a thin layer
through which the velocity changes continuously and rapidly
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 353
nuity, S0 , must be continuous across the surface. Hence
(10.7)
and discontinuities can only occur in the velocity components which have a
direction tangential to the surface S0 .
At some point on the surface of velocity discontinuity S0 , in figure 10.2(a),
let a local rectangular cartesian coordinate system be constructed with the
z axis coincident with the normal to the surface at the point. The velocities
as the surface is approached from region (1) and region (2) are v! 1>and v! 2>
and the tangential velocity components on the two sides of the surface, S0 ,
are vl 1 >and vf>, respectively. The tangential velocity discontinuity, vl 2>- vl 1 >,
is denoted V* and for convenience it is assumed that the direction of this
vector coincides with the x direction. The surface of velocity discontinuity
may be considered as the limiting position of a thin layer through which the
velocity changes continuously and rapidly from v< 1 >to v< 2>. For the situation
illustrated in figure 10.2(b) only the component in the x direction experiences
a rapid change, whilst the components in the other two directions are
essentially constant through the layer. It follows that the shear strain-rate
component, Yzx• is very much greater than the other strain-rate components
and, as the thickhess of the layer approaches zero, Yzx-+ oo. The shear stress
component of CT;j along the surface, S0 , in the x direction is denoted r.
Now consider an element of the surface dS0 . In region (1) the rate of work
done by the stresses is
- { crn Vn + 't"y vy + rvl 1 >} dS0
where rY and vY are the components of shear stress and tangential velocity
in the y direction.
The corresponding rate of work done by the stresses in the region (2) is
+ {crn vn + ry vy + rvl 2 >}dS0
and the algebraic sum of the rates of work done by the stresses acting on
the element of the surface, dS0 , is therefore
- 't" { vll) - vl2)} dSo = 't" { vl2) - vll>} dSo
= r[V*]dS0
Consequently, the rate of work done by the stresses acting on the surface
of velocity discontinuity, S0 , is given by

f r[V*JdS0
So
When more than one surface of velocity discontinuity exists in the body
the rate of work done by the stresses acting on the surfaces is

~ t r[V*JdS
354 ENGINEERING PLASTICITY

where the summation covers all the surfaces of velocity discontinuity


S = S 1 + S 2 + S 3 + .... For simplicity the summation sign can be omitted
and the virtual work equation then becomes

ti;v;dS+ fvX;v;dV= f/;jliijdV+ t't[V*]dS (10.8)

which is valid for any continuous medium in equilibrium. For the particular
case where O';j is an equilibrium plastic stress field and V; is the associated
velocity field

t 't[V*]dS = t k[V*]dS

where k is the yield shear stress.

10.4 UNIQUENESS THEOREM FOR A RIGID-PERFECTLY


PLASTIC BODY
The uniqueness theorem considered in this section is restricted to an isotropic,
incompressible, rigid-perfectly plastic material for which the yield criterion
is not influenced by the hydrostatic stress component.
Assume that a body of surface area, S, and volume, V, yields under the
action of surface tractions, J;, on the boundary surface of area, SF, and
prescribed velocities, V;, over the surface area Sv. The stress field must
satisfy the differential equations of equilibrium (10.1) at every point in the
body and the yield criterion must not be violated. The associated velocity
field must satisfy the condition of incompressibility, the velocity boundary
conditions and be related to the stress field according to the concept of the
plastic potential. The plastic potential was previously discussed in section
5.7.5 and it was shown that the tensor components of the plastic strain
increments during plastic flow are given by
de~= { og(u;)/ou;JdP = { of(u;)/ouij}dA.
where the yield criterion is assumed to exist for the material in the form
f( O';j) = 0 and dA. is a non-negative scalar constant of proportionality.
In a similar manner it can be shown that the tensor components of the
plastic strain-rate, l:;p are related to the stress field by
l:ij = i {of( O';j)/ouij} (10.9)
where i is a non-negative scalar constant of proportionality and the super-
script pis omitted since the material considered is rigid-perfectly plastic.
If the yield surface has edges or corners, as is the case for the Tresca
hexagon, then of(u;j)/ou;j is not uniquely defined. At the corners known as
singular points, eij is given by the generalisation of equation (10.9) due to
Koiter 6 which is
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 355
(10.10)

At a singular point on a yield surface a number of surfaces may meet and


f..( a;)= 0. The summation is taken over all the surfaces that meet at the
singular point. For an incompressible material, two surfaces meet at a
singular point. It should be appreciated that equations (10.9) and (10.10)
indicate the strain-rate vector representing, tij, corresponding to the plastic
stress state a;j, is normal to the yield surface.
Let a;j, V; and aG, v~ be two possible stress states and associated velocity
fields for the body which is yielding. For simplicity, assume that body forces
are absent and inertia forces are insignificant. The virtual work equation

t t
(10.8) then reduces to

T;v;dS= Iv ai\dV + T[V*]dS

Therefore t (T;- T~)(v;- vnds =IV (aij- aG)(Ilij- eG)dV

+ f
So
(k- T*)[ V*] dSD

+f s;
(k-T)[V*J*dSti (10.11)

where S0 and s:; are the surfaces of velocity discontinuity of the velocity
fields V; and v~, respective!~, [ V*] and [ V* ]* are the magnitudes of the
corresponding tangential velocity discontinuities and T and T* are the shear
stress components of the stress fields aij and aij, acting on the surfaces of
velocity discontinuity S0 and s:;, respectively.
The left-hand side of equation (10.11) is identically zero since T; = T~ on
the surface SF and v; = v~ on the surface Sv.
The principle of maximum work dissipation discussed in section 5.8 and
expressed by the inequality (5.93) is

Iv ( a;j- aij')li;j dV =Iv ( a;j- aij)A;j dV ~0


which, together with the inequality k ~ T, indicates that each term on the
right-hand side of equation (10.11) is non-negative and therefore each term
must be equal to zero. This implies that a;j = aij, except possibly where a
common rigid region of the velocity fields, V; and v~, exists and also where
both stress states, a;j and aij, occur on the same flat of a yield surface.
The uniqueness theorem for an isotropic, incompressible, rigid-perfectly
plastic material can thus be stated as follows :
356 ENGINEERING PLASTICITY

If two or more complete solutions can be determined for a problem then the
stress fields of the solutions are identical except possibly in the common non-
deforming regions of the solutions and when the stress states occur on the same
flat of a yield surface.

10.5 THE LOWER BOUND THEOREM


Let aii, eii, vi denote the actual solution to the problem of a rigid-perfectly
plastic body of total surface area, S, and volume, V, which is yielding under
prescribed surface tractions, J;, over the part surface, SF,
with the prescribed
velocity field, vp on the part surface Sv. The stresses and strain-rates are
related by the Levy-von Mises relations and satisfy the equations of equili-
brium and continuity.
For the actual stress distribution, aii, with velocity field, Vp the virtual
work equation (10.8) reduces to

I S
I;vidS= IV
aii8iidV +I k[V*]dS
So
0 (10.12)

when body forces are absent.


Now, let aij be any other statically admissible stress field which satisfies
the differential equations of equilibrium, the prescribed boundary conditions
on the part surface, SF,and nowhere violates the yield criterion. The stress
distribution, aij, may be discontinuous. For this statically admissible stress
field
(10.13)

where r* is the tangential component of the statically admissible stress


field, aij in the direction of the velocity discontinuity along the surfaces of
velocity discontinuity, S0 , of the actual velocity field vi. Now

Is I; vi dS =I I; vi dSF +I I; vi dSv
~ ~
(10.14)

Equation (10.12) can therefore be rewritten as

I I;vidSF+I
SF Sv
T;vidSv=I
V
aijf:ijdV+I k[V*]dS
S0
0 (10.15)

Also

(10.16)
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 357
since T; = Tt on the part surface SF.
Combining equations (10.13) and (10.16)

fSy
T{vidSv+f T;vidSF=J
SF V
u~eiidV+f
S0
t*[V*]dS0 (10.17)

Subtracting equation (10.17) from equation (10.15) gives

f Sy
(T;- Tt)vidSv = f (uii- u~)siidV + f (k- r*)[V*JdS
V So
0 (10.18)

The strain-rate vector, eii' is normal to the yield surface and the stress
vector, uii, is parallel to the strain-rate vector. However, in general, the
stress vector of the statically admissible stress field, u;j, will lie inside the
yield surface as shown in figure 10.3 and the stress difference vector (uii - u;j)
is defined by the dash line in figure 10.3. Because the yield surface is convex,
the angle subtended by the vectors (uii- u~) and tii is acute and therefore
their scalar product is positive, that is,
(10.19)
The equality sign in expression (10.19) is only applicable when the stress
fields, uii and uij, differ by an amount equal to a uniform hydrostatic pressure
since the yield criterion is unaffected.
The magnitude of the velocity discontinuity, V*, in the actual field is
unknown. However, since k ~ r* and k [ V*] > 0, the second integral on the
right-hand side of equation (10.18) is non-negative. It then follows that

f Sv
(T;- Tt)vidSv ~ 0

or (10.20)

Figure 10.3. Illustrating that the angle subtended by the stress vector difference
(uii-u~) and the strain-rate vector eii is acute for
a convex yield surface
358 ENGINEERING PLASTICITY

Expression (10.20) represents the first extremum principle which shows that
a lower bound for

can be obtained from a statically admissible stress field:

The rate ofwork done by the actual surface tractions with prescribed velocities
is greater than or equal to the rate of work done by the surface tractions corres-
ponding to any other statically admissible stress field.

10.6 THE UPPER BOUND THEOREM


As before, let crij, eij, vi constitute the solution to the problem. Together with
this actual state now consider another kinematically admissible velocity field,
vt, which satisfies the boundary conditions on the part surface, Sv, and also
the requirements for incompressibility of the material. The velocity field,
vt, may be discontinuous on certain surfaces, S0 .
It should be appreciated that in deriving the first extremum principle
the differential equations of equilibrium were required to be satisfied at all
points in the body. It was not essential that restrictions should be imposed
on the strain-rates as derived from the prescribed velocity fields. However,
in deriving the second extremum principle the approach is reversed. The
equations of equilibrium may not be satisfied and, hence, crij is not necessarily
statically admissible. Nevertheless, the velocity field must be kinematically
admissible.
The virtual work equation (10.8) is applicable to both the actual stress
distribution, crij, and the kinematically admissible velocity field, vt, that
satisfies the requirements for incompressibility of the body and the velocity
boundary conditions on the part surface, Sv, so that

f s
T,vtdS= f
v
uijeijdV + f r[V*]dS~
s;
(10.21)

If crij is a stress field not necessarily statically admissible which produces


the strain-rate field
eij = ( 1/2){ ( ov{/ox) + (ovj!oxJ} (10.22)

then f v ( crij - cri)eij d v~ 0


Therefore
f f
T, vt dS = crij eij d V + f" r [ V*] dS~
~f f
S V S0

crij eij d V + k [ V*] dS~ (10.23)


v s;
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 359
since k ~ r. On the part surface, Sv, v( =vi. Also

I ~v(dS =I
S Sv
~vidSv +I ~v(dSF
SF
(10.24)

Expression (10.25) is the second extremum principle which shows that an


upper bound for

which is the rate of work done by the unknown tractions, ~, on the part
surface, Sv, can be determined from a kinematically admissible velocity
field:

The rate ofwork done by the actual surface tractions with prescribed velocities
is less than or equal to the rate of work done by the surface tractions corres-
ponding to any other kinematically admissible velocity field.

If the mode of deformation is plane strain and can be envisaged as rigid


blocks of material which have a motion relative to each other at surfaces of
velocity discontinuity then Johnson 7 suggests that the first integral on the
right-hand side of expression (10.25)

Iv u~e~dV=O
Also, for many applications it is found that

I ~v(dSF
SF
=0

Thus, for these particular situations, expression (10.25) reduces to

I ~vi
Sv
dSv ~I * k[ V*] dS;';
So
(10.26)

An elementary justification for expression (10.26) for the upper bound


theorem applied to plane strain deformation involving simple patterns of
tangential velocity discontinuities occurring on surfaces separated by rigid
regions, is given in the next section.

10.6.1 The upper bound theorem applied to plane strain deformation


Consider a parallelepiped of material ABCD of vertical height, H, and of
360 ENGINEERING PLASTICITY

Line of velocity
discontinuity

(a) Physical diagram

(b) Hodograph lei Physical diagram 2

Figure 10.4. Assumed mode of plane strain deformation and hodograph for
determining the rate of internal energy dissipation for a rigid-perfectly plastic material

unit length in the z direction, shown in the physical diagram of figure 10.4(a)
to be moving to the left with unit velocity. It is assumed that the deformation
is plane strain in the (x, y) plane and the material is initially rigid. However,
on crossing the line XX, which is the trace of a plane surface of velocity
discontinuity, the parallelogram ABCD is distorted into the parallelogram
A'B'C'D' which moves with a velocity having, say, a magnitude, v2 , and a
direction which is at an angle, oc, to its original direction of motion. The
distortion of the parallelogram ABCD is due to the shear stress, -r, acting on
the surface of velocity discontinuity represented by the line XX.
Referring to the hodograph of figure 10.4(b) it will be seen that the vector
oa represents the initial unit velocity of the element which can be resolved
oe
into tWO COmponents, namely, the VeCtOr of magnitude VN, which is normal
ea
to the line XX, and the vector of magnitude va, which is parallel to the line
XX. Mter crossing the line XX the velocity of the distorted element is repre-
-+
sen ted by the vector ob of magnitude v2 , and direction at an angle oc to the
vector oa. Since the material is assumed to be incompressible, for constancy of
volume, the component of velocity normal to the line XX must remain
constant. The component of the vector Ob normal to the line XX must
therefore be given by the vector De of magnitude VN. The component of Ob
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 361
parallel to the line XX is then given by the vector e!> of magnitude vb. The
vector ab of magnitude vb - va = V* thus represents the tangential velocity
discontinuity occurring on the surface of velocity discontinuity. In reality,
the velocity discontinuity is not instantaneous, which would require the shear
strain-rate to be infinite, but is a rapid change in tangential velocity occurring
in a narrow region of the material.
In the physical diagram of figure 10.4(c) the parallelogram ABCD is shown
distorted into the parallelogram A'B'C'D' due to the shear stress, -r, producing
an angle of distortion, oc.
The energy dissipated by the shear stress, -r, during the distortion per unit
length in the z direction is given by
E = -r. BC. relative displacement CC'
The rate of internal energy dissipation per unit length in the z direction
dE/dt = -r .BC.(CC'/t)
where t is the time for the element to cross the line XX. Since the element
was initially moving towards the line XX with unit velocity, CD/t = 1 or
t = CD. Therefore
dEjdt = -r.BC.(CC'/CD)
The triangle CDC' with included angle oc, in the physical diagram of figure
10.4(c), is similar to the triangle aob in the hodograph of figure 10.4(b).
Therefore
CC'/CD = abjoa = V*/1
and dE/dt = -r.BC. V* (10.27)
When the material, assumed to be rigid-perfectly plastic, yields and
develops its maximum resistance to deformation, -r = k, the yield shear stress
in plane strain. Consequently, the rate of internal energy dissipation cannot
exceed ks V* where sis the appropriate length of the line of tangential velocity
discontinuity XX. The rate of work done by the external forces per unit
length in the z direction is
dWfdt ~ dE/dt = ksV* (10.28)
For a velocity field consisting of a number of straight lines of tangential
velocity discontinuity
dW/dt ~ IksV* (10.29)
If the line XX over which the velocity discontinuity occurs is curved, then
dS can be substituted for BC in equation (10.27) and expression (10.28)
becomes

dW jdt ~ dE/dt = l kV* dS (10.30)


362 ENGINEERING PLASTICITY

the integration being performed along the appropriate length of the line XX
where the magnitude of the tangential velocity discontinuity may be different
at each point.
It is usual to assume that the best upper bound estimate is provided by
the particular configuration of velocity field consisting of a pattern of
tangential velocity discontinuities which gives the least value of dE/dt.
However, since plastic energy dissipation is a non-conservative process the
minimum work principle may fail to produce a solution which agrees with the
actual problem. It is therefore preferable to consider that the assumed mode
of deformation which approaches the actual mode of deformation is likely to
give the best upper bound estimate of the required forming parameter,
irrespective of whether dEjdt is the least value for assumed configurations of
velocity field.
It should also be remembered that in the above consideration, body forces
such as gravitational attraction and inertia forces are assumed to be absent.
The effects of strain-hardening of the workpiece material are also neglected.
The study is therefore restricted essentially to the quasi-static, plane strain
deformation of a rigid-perfectly plastic material.

10.7 SOME APPLICATIONS OF THE UPPER BOUND


THEOREM TO PLANE STRAIN DEFORMATION
(JOHNSON UPPER BOUND ESTIMATES)
Before proceeding, it should be recalled that a slip line field solution to a
plane strain deformation problem is incomplete if no attempt is made to
extend the statically admissible stress field into the non-deforming regions.
Many solutions which have been proposed do not include an equilibrium
stress distribution satisfying the boundary conditions and ensuring that the
yield criterion is not violated which is extended into the assumed rigid regions.
In these circumstances, the slip line field solution does not satisfy the require-
ments of the lower bound theorem. However, the kinematically admissible
velocity field always satisfies the requirements of the upper bound theorem.
Incomplete slip line field solutions for plane strain deformation are therefore
strictly upper bound solutions. The relation between an upper bound solution
and the corresponding slip line field solution for a plane strain deformation
problem can now be more easily appreciated.
There are numerous examples of upper bound solutions to plane strain
metal forming problems which could be considered. It is only possible here to
present some examples which are intended to provide the reader with an
adequate understanding of the principles involved and the appropriate
background to appreciate the various applications which are to be found
in the published literature. Hopefully, the reader will gain the expertise
required to enable other upper bound estimates to be attempted for the
approximate solution to industrial metal forming problems which otherwise
may prove difficult.
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 363
10.7.1 Compression of a prismatic block between frictionless, rigid,
parallel platens (width/thickness ratio, WfT = 1)
One of the simplest examples of the application of the upper bound theorem
to initially consider is that of compression between smooth, rigid, parallel
platens. The slip line field solution of this problem for integral values of
width/thickness ratio was considered in section 7.4.6 and it was stated that
the uniformly distributed platen pressure can be shown to be equal to 2k,
where k is the yield shear stress of the deforming material.
For simplicity, it will be assumed that a rectangular section block of metal
of thickness 2T is compressed slowly between smooth, rigid, parallel platens
of width 2W, which approach each other with a relative velocity of 2, as
illustrated in the physical diagram of figure 10.5(a). The width/thickness
ratio for this case is therefore W jT = 1. As for the slip line field solution of
figure 7.36, the overhang of metal on each side of the platens is assumed to
be rigid.
Unit velocity
Line of tangential
velocity discontinuity
s = 2112 T

Unit velocity
2W

(a) Physical diagram

(b) Complete hodograph

Figure 10.5. Compression of a rectangular section block of metal between


smooth, rigid, parallel platens (width/thickness ratio, W /T = 1): (a) physical diagram;
(b) complete hodograph
364 ENGINEERING PLASTICITY

The intersecting lines or cross shown in the physical diagram represent


lines of tangential velocity discontinuity which intersect each other at
right-angles and meet the platens at angles of n/4. In the physical diagram,
it is convenient to designate the lines of tangential velocity discontinuity by
employing a notation such that the assumed rigid regions between these lines
are labelled using capital letters, in a manner similar to that when using
Bow's notation in relation to solving for the forces acting in the members of
a loaded space frame. Thus in the physical diagram of figure 10.5(a), four
lines of tangential velocity discontinuity can be designated as, namely, AB,
BC, CD and DA which are of equal length, s = 2 1 12 T. It is worth noting that
it is usual to designate stationary tooling, such as a die, and other stationary
regions using the capital letter 0.
The complete hodograph for the problem is given in figure 10.5(b), in
which corresponding lower case letters are used to designate the vectors
representing the velocity discontinuities in magnitude, direction and sense.
The upper platen moves downwards with unit velocity relative to earth. The
assumed rigid region, A, which is in contact with the upper platen, is therefore
constrained to move with the same velocity. In the hodograph, the velocity
of any particle in the region, A, is represented by the vector, oa, drawn to a
suitable scale to represent the magnitude of unity and to give the correct
direction and sense. All particles in the rigid region B are constrained to move
horizontally to the right. From the constancy of volume equation, it can be
easily shown that the magnitude of the velocity of all particles in the rigid
region B is unity. However, for the moment, it will be assumed that the
magnitude of this velocity is unknown. Nevertheless, the direction and sense
of the velocity is known. An indefinite vector can therefore be drawn from
the origin, o, in the hodograph which is horizontal in direction and where
the sense is from left to right. The point b will then be somewhere on this
line. In order that particles originally in the region A, which were moving
vertically downwards, eventually move horizontally to the right when they
enter region B it is necessary that their velocities are subjected to a sudden
change. This occurs as a tangential velocity discontinuity on the line of
tangential velocity discontinuity, AB, as the particles cross this line. The
vector representing this velocity discontinuity of magnitude, V* = 2 112 ,
is given by ab in the hodograph which is drawn from the point a, parallel
to the line of tangential velocity discontinuity AB, in the physical diagram.
Point b is thereby defined and consequently the magnitude of the tangential
velocity discontinuity, V*, is determined. Also, the magnitude of the velocity,
V, of all particles in the region B is then shown to be equal to unity. The
hodograph is completed by considering the other regions in a similar manner.
Let p be the uniformly distributed platen pressure, then from expression
(10.29)
dW/dt < 2p.2W.l ~ k~)V* = 4k.2 1i 2 T .2 1 12 = 8kT
or p ~ 2k = Y(Tresca) or 2 Y /3 112 (von Mises) (10.31)
where Y is the plane strain yield stress.
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 365
Because of the symmetry of the problem it is really only necessary to
consider one-quarter of the physical diagram and the corresponding quarter
of the hodograph of figure 10.5.
The upper bound solution given here for W /T = 1, with the particular
configuration of lines of tangential velocity discontinuity which intersect
each other at right-angles and meet the smooth platens at angles of n/4, can
be extended for other integral values of W /T. It has been shown by Slater 8
that also p ~ 2k for these cases.

10.7.2 A more general upper bound solution for the compression of a


prismatic block between smooth, rigid platens
The lines of tangential velocity discontinuity shown in the physical diagram
of figure 10.6(a) are assumed to meet the platens at equal angles, lJ, where (}
is not necessarily equal to n/4. One-quarter of the corresponding hodograph
is given in figure 10.6(b).
From the hodograph, the magnitude of the velocity discontinuity, V* =
cosec lJ, and the length ofthe lines of velocity discontinuity, such as AB = s =

Unit velocity

s• Wscc9/n

2W

Unit velocity
(a)

b d

:~~
a ---------c--

(b)

Figure 10.6. Compression of a rectangular section block of metal between


smooth, rigid, parallel platens: (a) physical diagram involving a pattern of straight line
tangential velocity discontinuities which meet the platens at equal angles (}; (b) one-
quarter of the corresponding hodograph
366 ENGINEERING PLASTICITY

1·07'r-----,..-----,..-- ---,..-----,..--------,

1-06

1·05 ~--Slip line field solution


(Green)
1·04
p/2k
1·03

1·5 2·0 2·5 3·0 3·5


WIT
Figure 10.7. Variation of pj2k with W/T for compression between smooth,
rigid, parallel platens

W sec e;n where n is the number of intersections of the lines of velocity


discontinuity on the horizontal axis of symmetry of the metal block.
From the physical diagram of figure 10.6(a), the relation between Wand
Tis found to beT= W tan ejn and the uniformly distributed platen pressure,
p, is given by
p. W.l:::;: k LsV* = kns cosec e = kn(W sec ejn) cosec e
Therefore pj2k:::;: 1/sin 2e:::;: (cote+ tan e)j2:::;: { (r/n) + (njr) }/2 (10.32)
where r = W jT.
The variation of pj2k with W/T is presented graphically in figure 10.7 and
is compared with the slip line field solution given by Green 9 • It will be seen
that the value of the platen pressure oscillates and that the upper bound
solution is exact for the minimum value which occurs for integer values of
W jT. For these cases the lines of tangential velocity discontinuity meet the
platens at angles of e = n/4 and when W /T = 1, figure 10.7 shows that,
pj2k = 1, which agrees with the solution previously obtained in section 10.7.1.

10.7.3 Indentation of a semi-infinite medium by a perfectly smooth,


flat, rigid punch
The slip line field solution for the indentation of a semi-infinite medium by
a flat, rigid punch was considered in section 7.4.3. Two solutions were
obtained corresponding to the indenter being partially rough and perfectly
smooth. Both solutions give a punch indentation pressure of p, where
pj2k = { 1 + (n/2)} ~ 2.57.
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 367
Unit velocity

0
a=60

(a)

(b)

Figure 10.8. Indentation of a semi-infinite medium by a perfectly smooth, flat,


rigid punch: (a) physical diagram showing the proposed pattern of straight lines of
tangential velocity discontinuity forming equilateral triangles; (b) the corresponding
half-hodograph

By referring to the hodographs of figures 7.28(b) and 7.28(c) for the two
solutions it can be deduced that, at incipient yield, the material beneath the
punch tends to move downwards and outwards which causes 'piling-up' at
the edges of the punch.
An approximation to the mode of deformation for a perfectly smooth
indenter is suggested in the physical diagram of figure 10.8(a) which consists
of a proposed pattern of straight lines of tangential velocity discontinuity.
Because of the symmetry ofthe system only the right-hand half of the physical
diagram need be considered. In this case, these straight lines of ~angential
velocity discontinuity are designated OA, AB, OB, BC and OC which,
together with the punch face and the stress free surface, form the sides of
three equilateral triangles of side length, s =a, where 2a is the width of the
indenter and the angles ct = 60°. The material below the pattern of tangential
368 ENGINEERING PLASTICITY

velocity discontinuities in the region, 0, is considered rigid and stationary.


The lines OA, OB and OC together form the boundary between the plasti-
cally deforming region and the non-deforming region.
All particles, such as P, initially in the region A, are constrained to move
vertically downwards with the same velocity as the indenter, D, that is, unit
velocity represented by the vector, od, in the corresponding half-hodograph
of figure 10.8(b). However, these particles also slide parallel to the line of
tangential velocity, OA, and parallel to the indenter face AD. The point, a,
in the hodograph is thus defined as the point of intersection of the two
indefinite vectors representing the correct direction and sense of the tangential
velocity discontinuity across OA and of the velocity of sliding of the region,
A, relative to the indenter, D, respectively. The lengths of the vectors oa and
da then define the magnitudes of these velocities, respectively, to the same
scale that the length of the vector od represents unit velocity.
As the particles cross the line of velocity discontinuity, AB, to move into
region B, they are subjected to a velocity discontinuity and sheared so as to
move parallel to the line of velocity discontinuity, OB. A further change in
the direction of motion of the particles occurs when they cross the line of
velocity discontinuity, BC, to move into region C, since they are also con-
strained to move parallel to the line of velocity discontinuity, OC. The
hodograph is therefore extended and completed by following a similar
procedure to that described for completing the triangle oda.
From the physical diagram, the length of the five equal lines of tangential
velocity discontinuity is given by OA = AB = OB = BC = OC = s =a. It
should be appreciated that the line AD is not a line of tangential velocity
discontinuity, since the indenter, D, is perfectly smooth and the region A
slides relative to the indenter and no internal energy is dissipated at the
interface, AD. Similarly, the stress free surface is not a line of tangential
velocity discontinuity.
From the hodograph, the magnitude of the equal velocity discontinuities
is oa = ab = ob = bc = oc = V* = 2/3 112 .
An upper bound estimate of the uniformly distributed indentation
pressure, p, is given by
dW/dt ~ dEjdt = l)sV*
Therefore p.a.l ~ k[(OA.oa) + (AB.ab) + (OB.ob) + (BC.bc)
+ (OC.oc)]
~ 5k(sV*)
~ 5k(a.2/3 112 )
~ lOka/3 112
or pj2k ~ 5/3 112 R; 2.89 (10.33)
If a pattern of tangential velocity discontinuities had been chosen such
that equal isosceles triangles were formed instead of equilateral triangles,
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 369
with the angles rx =I= 60°, then it can be shown that the minimum value of
pf2k = 2312 when rx =tan - 1 (2 112). Also p/2k = 3 when rx = n/4.

10.7.4 Drawing and extrusion through a symmetrical wedge-shaped die


Consider a moderate fractional reduction to be effected during the plane
strain sheet drawing through a perfectly smooth, symmetrical wedge-shaped
die of semi-angle, a, when the thickness of the sheet is reduced from 2H to 2h.
An acceptable upper bound estimate of the drawing stress, t, can be obtained
by assuming a pattern of straight line tangential velocity discontinuities
which form a single triangular field, shown as the outline in the physical
diagram of figure 10.9(a). Only the upper half of the physical diagram is given
since the system is symmetrical about the horizontal axis. The position of the

v
t

~ p
(a)

(b)

Figure 10.9. Drawing through a symmetrical wedge-shaped die with a moderate


fractional reduction: (a) physical diagram involving a pattern of straight line tangential
velocity discontinuities which form a single triangular field; (b) the corresponding
hodograph
370 ENGINEERING PLASTICITY

point, P, is arbitrary and can move along the horizontal axis of symmetry so
that the angle () varies. The rigid, stationary die is designated 0, and the
regions between the lines of tangential velocity discontinuity are A, B and C,
respectively. It follows that the lines of velocity discontinuity are designated
as AB and BC for the case when the die surface is perfectly smooth.
The undeformed material to the right of the line of velocity discontinuity,
AB, is assumed to be moving, as a rigid block, from right to left with unit
velocity such that steady state conditions prevail. Consequently, all particles,
such as Q, in region A, have unit velocity represented by the vector oa, drawn
to a suitable scale in the corresponding half-hodograph of figure 10.9(b).
As the drawing process proceeds, the particles cross the line of velocity
discontinuity, AB, to enter region Band are subjected to a tangential velocity
discontinuity across AB in order that they move parallel to the rigid die
surface, OB, at an angle,()(, to the horizontal axis of symmetry. The intersec-
tion of the two indefinite vectors correctly defining the direction and sense
of these velocities then defines the point b in the hodograph. The lengths of
the vectors ab and ob give the magnitude of the tangential velocity disconti-
nuity, A V~, across AB and the magnitude of the velocity of all particles in
region B relative to the stationary die surface to the same scale that the length
oa
of the vector represents unit velocity. When the particles in region B move
into region C they cross the line of velocity discontinuity, BC, and are
subjected to a tangential velocity discontinuity such that on entering region
C to become the drawn product they move, once again, parallel to the
horizontal axis of symmetry with a velocity V . These two indefinite vectors
which define the direction and sense of the tangential velocity discontinuity,
BC, and the velocity of the drawn product, V, intersect to define the point c
in the hodograph. The length of the vector, bC, gives the magnitude of the
tangential velocity discontinuity, 8 V~, across BC and the length of the
vector, oc, gives the magnitude of the velocity of the drawn product. If the
material is assumed to be incompressible, then for constancy of volume,
Vh = l.H or V = Hjh.
Let t be the drawing stress required. An upper bound estimate of the
drawing stress is given by

d W /dt ~ dEjdt = L)s V*


Therefore t.h. V = k[(AB.ab) + (BC. be)] (10.34)
when the die surface is perfectly smooth, there being no dissipation of internal
energy along the die surface OB. However, if the die were perfectly rough
instead of being smooth then an additional term would have to be included
on the right-hand side of equation (10.34) for the dissipation of energy
occurring at the die surface. In this case, equation (10.34) becomes
t.h. V = k[(AB.ab) + (BC.bc) + (OB.oo)] (10.35)
In general t.h. V = L)sV* or t.H = kl:sV*
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 371
Hence, the dimensionless drawing stress can be stated as
t/2k = '[,sV*/2H (10.36)
By measuring the lengths of the lines of tangential velocity discontinuity,
s, in the physical diagram of figure 10.9(a) and the lengths of the vectors in
the corresponding hodograph of figure 10.9(b) representing the magnitudes
of the velocity discontinuities, V*, the value of the drawing stress, t, can be
determined using equation (10.36), where k is the yield shear stress of the
deforming material. For the case considered in figure 10.9 which is shown in
outline and where ex= 40°, (} = 45° and the fractional reduction R =
(H- h)/H ~ 0.56, the values can be tabulated as given in table 10.1. The
physical diagram is drawn to a scale, H = 10 em, and the hodograph to a
scale, 5 em= unit velocity. Using equation (10.36)
t/2k = '[,sV*/2H = 2.031H/2H ~ 1.016 (10.37)
The slip line field solution for the corresponding configuration gives a value
of t/2k = 0.95.
Equation (10.36) therefore produces a certain value of t/2k depending on
the position of the point P along the horizontal axis of symmetry or the
chosen angle fJ. Assuming that the best upper bound estimate for tj2k is
given by the least value, then various values of fJ can be chosen such as fJ 1 ,
and fJ 2 , corresponding with the point P moving to the alternative positions
ofP 1 and P 2 , respectively. The patterns of straight lines of tangential velocity
discontinuity for these alternative configurations are shown in the physical
diagram of figure 10.9(a) as short chain lines. Similarly, the alternative
corresponding hodographs are shown as short chain lines in figure 10.9(b).
It may be noted that as the point P moves along the horizontal axis of
symmetry in the physical diagram, so does the point b in the hodograph
move along the vector, Ob, produced if necessary. Values of the correspond-
ing dimensionless drawing stress, t/2k, can be determined as previously
indicated and the values plotted to produce the graph presented in figure
10.10, from which the minimum value oftj2k can be obtained corresponding
with the optimum value of fJ. It will be noted that the best upper bound
estimate for t/2k ~ 1.0 is given when the optimum angle(}~ 40°.

Table 10.1 Measurements from the physical diagram of


figure 10.9(a) andthehodographoffigure 10.9(b)
Line of velocity Tangential velocity
discontinuity Length,s discontinuity, V* sV*

AB 1.025H '!1> = 5.4/5 = 1.08 1.101H


BC 0.62H be= 1.45/5 = 1.49 0.924H

~sV* 2.031H
372 ENGINEERING PLASTICITY

2·2.---------,,--------,,--------,,-------,

1·0
:Optimum value ot9•40
0•820 30 60

Figure 10.10. Variation of the upper bound estimate for the dimensionless
drawing stress, t/2k, with the angle, 0, for drawing through a perfectly smooth and a
perfectly rough symmetrical wedge-shaped die

The corresponding curve for the case when the die is perfectly rough is
also presented in figure 10.10. It is then found that the minimum value of
t/2k ~ 1.62 corresponding with an optimum value of(}~ 32°. The slip line
field solution for this case gives t/2k = 1.48.
In this section, plane strain drawing through a symmetrical wedge-shaped
die has been considered. However, the same solutions are valid for the plane
strain forward extrusion of sheet material through a symmetrical wedge-
shaped die having the same die semi-angle, ex, and for the same fractional
reduction, R, if it can be assumed that the container is perfectly smooth. The
dimensionless drawing stress, t/2k, is of course replaced by the dimensionless
extrusion pressure, Pe/2k.
For large fractional reductions, a better upper bound estimate for t/2k
may be obtained by introducing a pattern of straight line tangential velocity
discontinuities consisting of more than one triangle as, for example, shown
in the physical diagram of figure lO.ll(a). The corresponding half-hodograph
is presented in figure lO.ll(b). An acceptable upper bound estimate of t/2k
for small fractional reductions can be obtained by choosing a velocity field
of the type shown in the physical diagram of figure 10.12(a) for which the
corresponding half-hodograph is that given in figure 10.12(b).

10.7.5 Forward extrusion through a square face die


The square face die may be considered as the particular case of a symmetrical
wedge-shaped die discussed in the previous section when the die semi-angle
ex= n/2. In this section, two cases will be considered:
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 373

-
Unit velocity

A
H

v
t

(a)

e
~ v-H~
d a I
1
o

~ c

(b)

Figure 10.11. An upper bound solution for drawing through a symmetrical


wedge-shaped die with a large fractional reduction: (a) physical diagram involving a
pattern of straight line tangential velocity discontinuities forming two triangles; (b) the
corresponding half-hodograph

(1) Where the die and container are smooth such that there is free flow of
material along the die face and there is no internal energy dissipation at the
die and container surfaces.
(2) Where the container is assumed to be smooth but a dead metal zone
(D.M.Z.) is formed at the die face such that the deforming material yields in
shear at the interface with the dead metal zone and internal energy is therefore
dissipated at this interface.

Assuming the die face to be smooth and the velocity field to consist of the
single triangle of straight line tangential velocity discontinuities shown in
the physical diagram of figure 10.13(a), then the corresponding half-
hodograph is that given in figure 10.13(b). Note that the die-material
interface, OB, is smooth and is not a surface of tangential velocity disconti-
nuity. The point, P, in the physical diagram of figure 10.13(a) can move along
the horizontal axis of symmetry and hence the angle, (}, is the independent
variable. From the physical diagram of figure 10.13(a):
z = h cot(}
374 ENGINEERING PLASTICITY

(BC) 2 = z 2 + h2 = h2 cotl 0 + h2 = h2 cot 2 0(1 + tan 2 0)


Therefore BC = h cot 0(1 + tan 2 0) 1 ' 2
(AB) 2 = z 2 + H 2 = h2 cot 2 0 + H 2 = h2 cotl 0(1 + x 2 tan 2 0)
where x = H /h. Therefore
AB = h cot 0( 1 + x 2 tan 2 0) 112
From the hodograph of figure 10.13(b):
ob = xtan 0
(ab) 2 = 1 + x 2 tan 2 0
Therefore ab =(1 + x 2 tan 2 0) 1' 2
and (bc) 2 = (ocf + (ob) 2 = x 2 + x 2 tan 2 0 = x 2 (1 + tan 2 0)
Therefore be = x( 1 + tan 2 0) 112

-
Unit velocity

A H
v

(a)

V-= H/h I

(b)

Figure 10.12. An upper bound solution for drawing through- a symmetrical


wedge-shaped die with a small fractional reduction: (a) physical diagram involving an
arbitrary pattern of straight line tangential velocity discontinuities; (b) the corres-
ponding half-hodograph
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 375

(a)

V=H/h•x

(b)

Figure 10.13. Forward extrusion through a smooth square face die: (a) physical
diagram involving a pattern of tangential velocity discontinuities consisting of a single
triangle; (b) the corresponding half-hodograph

Let P. be the extrusion pressure then since


d W /dt ~ dEjdt = ~)s V*

P•. H .1 = k[(AB.ab) + (Bc.bc)J


= k [h cot 0(1 + x 2 tan 2 0) 1 12 (1 + x 2 tan 2 0) 1 12
+ h cot 0(1 + tan 2 0) 1 i 2 x(1 + tan 2 0) 112 ]
= khcot 0[1 + x 2 tan 2 e + x(1 + tan 2 0)]
= kh cot0[1 + x + (x + 1)x tan 2 OJ
=kh(x+ 1)cot0[1 +xtan 2 0]
The extrusion pressure is therefore given by
P. = k{ (x + 1)/x} cot 8(1 + xtan 2 0) (10.38)
or P.f2k = { (x + 1)/2x} cot 8(1 + xtan 2 8) (10.39)
dp.fd8 = k { (x + 1)/x}{ sec 2 0(1 - x tan 2 O)jtan 2 0}
376 ENGINEERING PLASTICITY

For P. to have a minimum value, dp./dO = 0. Therefore


1- xtan 2 0 = 0
or tan 2 0 = 1/x
Hence tan 0 = (1/x) 112
and 0 =tan - 1 (1/x) 112 (10.40)
Combining equations (10.38) and (10.40) yields the minimum value of the
extrusion pressure
Pe(min) = k{(x + 1)/x}2x 112 (10.41)
and the minimum value of the dimensionless extrusion pressure is
Pe(min/2k = (x + 1)/x 112 (10.42)
The upper bound estimate of the dimensionless extrusion pressure, P./2k,
given by equation (10.39), obtained by assuming the single triangle pattern
of straight line tangential velocity discontinuities, is found to overestimate
the slip line field solution for the problem. This overestimate varies from
about 20 to 30 per cent, being greatest for large fractional reductions.
For a moderate fractional reduction it is known that a dead metal zone
may be formed which covers the entire die face. In effect, this converts the
square face die into a perfectly rough curved die. However, as an .approxi-
mation to the mode of deformation, it may be assumed that the formation
of a dead metal zone simply produces a perfectly rough, wedge-shaped die
of unspecified die semi-angle, r:t., as shown in the physical diagram of figure
10.14.
Assuming the fractional reduction to be known then, by following the
procedure adopted in section 10.7.4, the minimum value of the dimensionless
extrusion pressure, P.l2k, can be determined for values of the angle 0,

Unit .
-+--+-velOCity
I
I

Figure 10.14. Forward extrusion through a rough square face die with a dead
metal zone (D.M.Z.) present covering the entire die face. The physical diagram involves
a single triangular velocity field
EXTREMUM PRINCIPLFS FOR A RIGID-PERFECTLY PLASTIC MATERIAL 377
and for a range of values of the angle ex. The least of these minima may then
be regarded as the best upper bound estimate of the extrusion pressure for a
square face die with a dead metal zone present appropriate to that particular
fractional reduction.

(a) (b)

Figure 10.15. Other patterns of tangential velocity discontinuities for forward


extrusion through square face dies with large fractional reductions when a dead metal
zone (D.M.Z.) is present

(a)

(b)
Figure 10.16. (a) A multiple triangular pattern of tangential velocity dis-
continuities for forward extrusion through a square face die with a fractional
reduction R = 1/2 when a dead metal zone (D.M.Z.) covers the entire die face; (b) the
corresponding half-hodograph
378 ENGINEERING PLASTICITY

For large fractional reductions, a more accurate estimate of the extrusion


pressure is likely to be obtained by assuming other velocity fields of the
types shown in the physical diagrams of figure 10.15. Experimental evidence
suggests that, for very large fractional reductions, a dead metal zone is
formed in the corner at the junction between the die face and the container
surface. This dead metal zone does not extend over the entire die face.
However, the deforming material also shears along the lower part of the
die face. The velocity field proposed in figure 10.15(b) is therefore in accord
with the evidence available.
A kinematically admissible pattern of straight line tangential velocity
discontinuities for plane strain forward extrusion through a square face die,
where a dead metal zone is present covering the entire face of the die and the
fractional reduction R = 1/2, is presented in the physical diagram of figure
10.16(a). The velocity field is of the triangular type and, for the case consider-
ed, four similar triangles are involved. However, it will be appreciated that
the proposed velocity field could have involved any number of similar
triangles. The greater the number of triangles the more accurate is the upper
bound estimate of the extrusion pressure likely to be. The corresponding
half-hodograph is given in figure 10.16(b).
From the physical diagram of figure 10.16(a) OB = BC =CD= DE= EF
and AB = AC =AD= AE. From the hodograph of figure 10.16(b):
ab = ac = ad = ae
ob=ef
od ~=~=&
Since dW /dt ~ dE/dt = LksV*
p•. H.1 = k[(OB.ob) + (BC. bc) + (CD.cd) + (DE.de) + (EF.et)
+ (AB.ab) + (AC.ac) + (AD.ad) + (AE.ae)] (10.43)
where Pe is the extrusion pressure.
Because of the equal lengths of the straight lines of tangential velocity
discontinuities in the physical diagram of ·figure 10.16(a) and the equal
magnitudes ofthe vectors in the hodograph of figure 10.16(b), equation (10.43)
may be rewritten as
p•. H.1 = k[(OB.ob) + BC(bC + Cd +de)+ ab(AB + AC +AD+ AE)
+ (EF.cl)J
= k[(OB.Ob) + (BC.Lbc) + (ab.LAB) + (EF.enJ (10.44)
If the values are calculated and substituted into equation (10.44) it will
be found that the dimensionless extrusion pressure is given by

P./2k R:i 1.3 (10.45)


In the limit when the number of triangles in the velocity field becomes
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 379

(a)

v- H/h•2

(b)

Figure 10.17. (a) A pattern of tangential velocity discontinuities involving a


circular arc for forward extrusion through a square face die with a fractional reduction
R = 1/2 when a dead metal zone (D.M.Z.) covers the entire die face (b) the corresponding
half-hodograph

infmite, a circular arc is formed. The pattern of tangential velocity disconti-


nuities then becomes that shown in the physical diagram of figure 10.17(a)
and the corresponding half-hodograph is that given in figure 10.17(b). From
the physical diagram of figure 10.17(a) OB = BC = 2 1/2 and circular arc
AB = 2 112 .(n/2) = n/2 112. From the hodograph of figure 10.17(b):
- - - - 1/2
ob 1 = b 2c = ab 1 = ab 2 = 1/2
and circular arc b 1b 2 = (1/2 1i 2)(n/2) = nj2 312
Following the format of equation (10.44)
Pe·H .1 = k[(OB.ob 1) + (BC.arc b 1b 2) +(arc ab 1.AB) + (BC.b 2c)]
= k [ { 2 112. ( 1j 2)112} + {2 112. (n/2 3/2)} + { (1j 2)112. (n/ 2 112)}
+ {21/2 .(1/2)1/2}]
= k[1 + (n/2) + (n/2) + 1]
= k[n + 2]
Therefore the dimensionless extrusion pressure is
Pe/2k = (n + 2)/2H = (n + 2)/4 R; 1.29 (10.46)
380 ENGINEERI NG PLASTICITY

v
Y velocity _ . L __ _ __

(a)

v- IH-yl/lh-y l

(b)

Figure 10.18. Forward extrusion of a short billet through a square face die with a
dead metal zone (D.M.Z.) covering the entire die face: (a) physical diagram involving a
proposed pattern of straight line tangential velocity discontinuities; (b) the correspond-
ing half-hodograph

which agrees with the slip line field solution of this problem given by equation
(7.61) derived in section 7.4.1. The hodograph for the slip line field solution
presented in figure 7.25(b) and that of figure 10.17(b) are identical. This is
not surprising since it will be recalled that, in the slip line field solution,
tangential velocity discontinuities were assumed to occur at the boundary
between the plastically deforming region and the rigid regions.
A possible mode of deformation for the plane strain forward extrusion
of a short billet through a square face die is defined by the pattern of straight
line tangential velocity discontinuities shown in figure 10.18(a). The situation
depicted is the unsteady state phase which occurs at the end of the extrusion
process. It is assumed that the face of the punch or pressure pad A is smooth.
The corresponding half-hodograph is given in figure 10.18(b). For constancy
of volume
H .1 = (y .1) + {(h- y) V}
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 381
Therefore V = (H - y)j(h - y) (10.47)
Let H - h = d and h - y = z.
From the physical diagram of figure 10.18(a):
OB = (d 2 + 12)1' 2, BC = (z 2 + 12 ) 112 , sin(}= d/(d 2 + 12)1' 2 and
sin oc = zj(z 2 + 12)1' 2
From the hodograph of figure 10.18(b):
ab =tan(}= djl, ob = ab/sin (} = (d 2 + 12)112jl
and be= ab/sinoc = d(z 2 + 12)1' 2jzl
dWjdt ~ dE/dt = :EksV*
Therefore dW/dt = k[(OB.ob) + (BC.bc)]
= k[ { (d2 + l2)112(d2 + [2)1/2 /l}
+ {(z2 + [2)1/2. d(z2 + [2)1/2 / zl}]
= (k/l)[d 2 + 12 + {d(z 2 + 12)/z}]
= (k/l)[d 2 + 12 + d{z + (1 2/z)}] (10.48)
dW /dz = (k/l)[d {1 - Wjz2)}] (10.49)
If the extrusion pressure is to have a minimum value then d Wjdz = 0.
Therefore
1- W/z 2 ) =0
or [2 jz2 =1
and z=l
It then follows that h-y=z=l
and y=h-l (10.50)
When the point P is not coincident with the point Q on the horizontal
axis of symmetry, the velocity of the extruded product Vis such that the
billet ceases to maintain contact with the punch face or pressure pad over
the length PQ. The proposed model is therefore able to offer an explanation
for the formation of a cavity at the rear face of the billet leading to the extru-
sion defect, referred to as piping, which was previously discussed in section
8.5.2. The length, 2y, is related to the dimension of the cavity.

10.7.6 Coulomb friction present at workpiece-tooling interfaces


In the examples of the application of the upper bound theorem discussed
in section 10.7, the effect of Coulomb friction which may be present at an
interface between the workpiece and tooling was not considered. It will have
been noted that the conditions existing at such interfaces were assumed to be
either perfectly smooth, that is, frictionless or perfectly rough when the
382 ENGINEERING PLASTICITY

frictional stress attained the limiting value of the yield shear stress, k, of
the deforming material. The method employed does not permit the introduc-
tion of Coulomb friction directly since the pressure normal to a partially
rough interface when Coulomb friction is present is not readily calculable.
However, there are two simple methods which are used in an attempt to
overcome this difficulty. In the first method, the required forming parameter
is initially determined for the frictionless case as, for example, for the plane
strain drawing through a symmetrical, smooth, wedge-shaped die shown in
figure 10.9(a). The frictionless drawing stress, t, is calculated from which
the corresponding normal die pressure, q, may then be deduced. Assuming
that there is no variation of the normal die pressure due to the presence of
Coulomb friction then the drawing stress can be recalculated. In this parti-
cular example the drawing stress with Coulomb friction present at the
workpiece-die interface is (1 + J.LCOtoc) times that for the frictionless case,
where oc is the die semi-angle and J.l is the constant coefficient of friction.
It is not thereby implied that the solution obtained by this method constitutes
an upper bound for the drawing stress.
In the second method a coefficient, m, is introduced such that at a partially
rough interface the constant frictional stress is expressed as, mk, and this
value is substituted for k in the calculations. The normal procedure outlined
in section 10.7 is otherwise followed. The coefficient, m, can be regarded
as analogous to the coefficient of friction, J.l, and for all possible cases has
values 0 ~ m ~ 1, since when m = 0 the interface is smooth and if m = 1
the limiting condition of stiction or shearing friction exists. It will be recalled
that the same method was adopted in section 7.5.2. Plane strain compression
was considered with the limiting condition of stiction existing at the work-
piece-platen interfaces.
These two methods of accounting for the effect of Coulomb friction must
necessarily be approximate because, for many cases, the coefficient of friction,
J.L, or the coefficient, m, is likely to vary from point to point along the interface
such that there is a non-uniform distribution of normal die pressure. Never-
theless, constant values for J.L, m and k are often used even though these
assumed conditions may not be realised due to, for example, the strain
hardening characteristics of the deforming material, the effects of temperature
distribution and strain rate.

10.8 UPPER BOUND ESTIMATES FOR ANISOTROPIC METALS


The method of determining an upper bound estimate of the required forming
parameter for plane strain deformation of an isotropic, rigid, perfectly
plastic material involving a pattern of tangential velocity discontinuities was
demonstrated in section 10.7. This method can be extended so as to be
applicable to anisotropic metals, for which the yield criterion may be express-
ed by the anisotropic yield criterion of equation (4.42) proposed by Hill.
An upper bound estimate for the required forming parameter is determined
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 383
in a similar manner to that used in the previous section by obtaining the
summation l: ks V* for the lines of velocity discontinuity in the proposed
pattern. However, for the case of an anisotropic metal, account is taken of
the variation of the yield shear stress, k, with direction. To assist this operation
a polar diagram can be constructed to represent the variation of the yield
shear stress with direction when the anisotropic parameters for the material
have been established. In the paper by Johnson, de Malherbe and Venter 10
such a polar diagram is constructed and examples of its use are given.

10.9 LINES OF THERMAL DISCONTINUITY OR 'HEAT LINES'


AND TEMPERATURE JUMPS DURING A FAST METAL
FORMING PROCESS
The external work done during a metal forming process is nearly all dissi-
pated as heat and, in a fast process which may be considered as adiabatic,
the consequences of heat transfer by radiation and conduction become
insignificant.
It has already been postulated that, in a plane strain plastic deformation,
relative movement only occurs at the planes across which there is a tangential
velocity discontinuity and where internal energy is dissipated. Local heating
could therefore be expected to occur at these planes where a sudden increase
in temperature or temperature jump results in some way related to the
magnitude of the velocity discontinuity.
To imagine so-called lines of thermal discontinuity is thus a natural deve-
lopment. It can be expected that distinct regions or lines will exist where
a temperature jump occurs. These lines of thermal discontinuity or heat lines
may be visible during the process depending on the magnitude of the tempera-
ture jump and also the contrast between the heat lines and the surrounding
body of material.
An expression for the temperature jump, ll.(), which occurs at a single
line of thermal discontinuity corresponding to a single straight line of
tangential velocity discontinuity, such as XX in the physical diagram 1 of
figure 10.4(a), may be deduced as follows:

During a fast plastic deformation which may be considered as adiabatic


the internal energy dissipated in unit time is equal to the heat generated.
Hence, for a single line ofthermal discontinuity
ks V* = J x mass displaced in unit time x specific heat x temperature jump
= J .p(l.H .l).c.ll.(J
where p is the density of the deforming material, c is the specific heat and
J is the mechanical equivalent of heat. Therefore
ll.() = ks V* I J pcH = (ksl J pc)(V* IH) = (ksl J pc)(V* Is. sin ljJ)
since H = s. sin </J.
384 ENGINEERING PLASTICITY

From the hodograph of figure 10.4(b):


sin c/J = vN/1
Therefore AfJ = (k/ J pc)(V* /vN) (10.51)
and if the von Mises yield criterion is assumed to be applicable, k = Y /3 1' 2 ,
where Y is the plane strain yield stress and equation (10.51) becomes
(10.52)
The first term in equations (10.51) and (10.52) is peculiar to the deforming
material and is a property of that particular material. The second term,
which is obtained from the hodograph, is peculiar to the mode of deformation
and is simply a measure of the shear strain. In reality, no precise line of
thermal discontinuity can exist but a thin diffused region can appear across
which a temperature jump above the temperature of the surrounding material
will occur.

10.9.1 Temperature jumps during plane strain compression between


smooth, rigid, parallel platens (width/thickness ratio, W/T= 1)
The physical diagram and the corresponding complete hodograph presented
in figures 10.5(a) and 10.5(b), respectively, are appropriate to this problem
and the reader is referred to these. The straight lines of tangential velocity
discontinuity AB, BC, CD and DA of equal length shown in the physical
diagram may also be considered as lines of thermal discontinuity across
which temperature jumps, AfJ, occur.
Assume the workpiece material to be mild steel initially at a temperature
of 680°C, that is, below the normal forging temperature of about 1400°C and
for which the density p = 7.8 gjcm 3 , the specific heat c = 620 jjkgoC and the
yield stress, Y, at the temperature of 680°C and appropriate strain rate is
350 MPa. The mechanical equivalent of heat J = 1 mN/Joule in the S.l.
system of units.
Y /3 112 J pc = 350 X 106 /3 112 X 1 X 7.8 X 106 X 620 ~ 42°C
From the hodograph of figure 10.5(b):
ab = 6C = cd = da = V* = 2 1' 2
and (vN)AB = (vN)BC = (vN)CD = (vN)DA = (1/2) 112
Therefore V* fvN = 2 1' 2 /(1/2) 1 ' 2 = 2
It should be noted that the vector, vN, which is the component of the
particle velocity normal to the line of tangential velocity discontinuity or
thermal discontinuity is given by drawing a line from the origin, o, of the
hodograph normal to the line of tangential velocity discontinuity considered,
which is produced if necessary.
The temperature jump which occurs across each of the lines of thermal
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 385
discontinuity AB, BC, CD and DA is given by equation (10.52). Therefore
AO = (Y /3 112 J pc)(V* jvN) ~ 42 x 2 ~ 84°C (10.53)
The existence of lines of thermal discontinuity or 'heat lines' were appa-
rently first reported by Massey 11 in 1921, although he was unable to provide
photographic or other evidence of his observations. However, in 1964, an
experiment was conducted by Johnson, Baraya and Slater 12 to verify the
prediction of equation (10.53). A prismatic block of mild steel was preheated
to a temperature of 680°C and forged rapidly between the flat dies of an
industrial power hammer with an impact velocity of about 4 mjs. A thermal
cross was evident corresponding to the lines of thermal discontinuity which
was photographed using high speed film. By means of an optical pyrometer
the temperature rise was measured which was found to be about 100°C.
The minor difference between this temperature jump and that predicted
by equation (10.53) is easily accounted for by such factors as the geometry
of the actual situation differing from that assumed, frictional resistance at
the interface between the deforming block and the power hammer dies,
and inaccuracy of the assumed material properties, etc.
The important significance of this experiment and of the concept intro-
duced in section 10.9 is that the physical existance oflines of velocity disconti-
nuity is demonstrated. Furthermore, the existence of lines of thermal
discontinuity is important in plastic deformation processes because the
temperature that may be actually attained locally in the workpiece material
can thus exceed that anticipated. This phenomenon can lead to an overheated
structure or an unexpected metallurgical phase change. In certain metals,
such as air-hardening steels, it is also possible for internal martensitic
structures to be developed during rapid cooling after an impact process that
can lead to catastrophic failure which otherwise may be difficult to account
for. These phenomena are discussed at greater length by Slater 8 •

10.9.2 Temperature distribution during a fast metal forming process


Nearly all the plastic work done during a metal forming process is eventually
dissipated as heat. It is therefore inevitable that a temperature distribution
will exist throughout the deforming material and, in some instances, the
consequences may be of considerable metallurgical importance.
In such cases, an estimation of the temperature distribution can prove
valuable. For fast plane strain metal deformation this is a relatively simple
problem because heat transfer due to radiation and conduction may be
considered as insignificant. It has been shown by Tanner and Johnson 13
that for some engineering metals including copper, aluminium and mild steel
heat by thermal conduction does not propagate upstream at relatively low
forming speeds. Thermal conduction effects are neglected in the example to
be considered.
For simplicity, the method will be described by reference to forward
386 ENGINEERING PLASTICITY

Ia I (b)

d
I• V= I +2sina•2
a I
•j
o

~ c2
bc 1

lei

(d)

Figure 10.19. (a) Top half of slip line field for extrusion through a smooth wedge-
shaped die of semi-angle ex= 30° with a fractional reduction R = 2sincx/(1 + 2sincx)
= 1/2; (b) pattern of tangential velocity discontinuities to replace the slip line field of
(a) where a chord is substituted for a circular arc; (c) half-hodograph corresponding
with the slip line field of (a); (d) half-hodograph corresponding with the pattern of
tangential velocity discontinuities of (b)

extrusion through a smooth wedge-shaped die of semi-angle, a = 30°, with a


fractional reduction R = 2 sin a/(1 + 2 sin a)= 1/2. The slip line field which
is valid for this particular fractional reduction is given in figure 10.19(a) and
consists of an isosceles triangle, EFK, and a circular sector, FGK, which
subtends an angle, a, at its centre. Radial slip lines such as GF and KF are
a lines and GKE is a f3 line.
The first step in the method is to approximate the slip line field by repiacing
circular arcs with chords or sectors with an appropriate number of triangles.
The degree of accuracy required determines the system of chords or the
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 387
number of triangles to be used. For the case considered here it is only neces-
sary to replace the circular arc, GK, by a single chord, that is, replace the
circular sector, FGK, by a single triangle. The slip line field of figure 10.19(a)
is thus replaced by the triangular pattern of tangential velocity discontinuities
shown in figure 10.19(b). This kinematically admissible velocity field can be
used to determine an upper bound estimate for the extrusion pressure. It can
be shown that the slip line field solution gives the dimensionless extrusion
pressure, P.l2k = (1 + a)R ~ 0.762 and the corresponding upper bound
estimate, which is derived later, is P./2k ~ 0.767 using the pattern of tangen-
tial velocity discontinuities of figure 10.19(b).
Secondly, the half-hodograph corresponding with the physical diagram of
figure 10.19(b) is constructed. This is given in figure 10.19(d) and, for com-
parison, the half-hodograph for the slip line field solution of figure 10.19(a)
is presented in figure 10.19(c). The hodograph enables the velocity of any
particle during the extrusion to be determined.
Thirdly, a number of stream tubes are defined by horizontal lines which
meet the entry boundary line of the velocity field at points of discontinuity.
There are only two stream tubes, I and II, for this particular problem and
only one discontinuity in the entry boundary line defined by the point P in
the physical diagram of figure 10.20(a). The directions of the stream lines of
particles which pass through these points, such as P, are obtained from the
hodograph. For example, the direction of the stream line for all particles in
regions cl and c2 is given by the direction of the vector, ocl or oc2' shown as
a short chain line in the half-hodograph of figure 10.20(b). Particles moving
within a given stream tube have the same history of deformation.
Consider stream tube I in figure 10.20(a), where it is assumed that the
material approaches the line of tangential velocity discontinuity or thermal
discontinuity, AB, with unit velocity. If there is no heat transfer such that
the extrusion process may be considered as adiabatic then the temperature
jump which occurs across the line, AB, is given by equation (10.52) that is
A(}AB = ( y /3112 J pc)(V* /vN )AB
which can be rewritten as
M} AB = b( V* /vN )AB
where b=(Y/3 112 Jpc) is a constant for the deforming material and the
quotient (V*jvN)AB for the line of thermal discontinuity, AB, is obtained
from the hodograph of figure 10.20(b).
The material in region B is rigid and remains at constant temperature
until it crosses the line of thermal discontinuity, BC 1 • Then, for a second
time, it is subjected to a temperature jump given by
A(}Bc 1 = b(V*/vN)Bcl
A further temperature jump occurs as the material crosses the line of thermal
discontinuity, C 1D 1, to become the extruded product. This temperature
388 ENGINEERING PLASTICITY

(a)

(b)

Figure 10.20. (a) Physical diagram showing stream tubes and relative temperature
distribution following figure 10.19 (b) and (d); (b) corresponding half-hodograph
indicating thtt construction for the determination of the value of, vN, for each line of
thermal discontinuity

jump is given by
.::\Oc,o, = b( V*/vN)c,o,
The final temperature of the extruded product in the region D 1 , above that
of the undeformed material in region A, is the sum of the separate temperature
jumps
b {( V*/vN)AB + (V* /vN)Bc 1 + ( V*/vN)c,o,}
Similar calculations are made for stream tube II to determine the tempera-
ture jumps as the material in this stream tube crosses the lines of thermal
discontinuity, AC 2 and C 2 D 2 . It will be observed that in stream tube I,
material crosses the three lines of thermal discontinuity AB, BC 1 and C 1 D 1
whereas the material in stream tube II crosses only two lines of thermal
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 389
discontinuity, AC 2 and C 2 D 2 • The temperature of the extruded product on
the outside in region D 1 could therefore be expected to be different to that
on the inside in region D 2 •
If the temperature of the undeformed material in region A is known and
the material constant {) = Y /3 1' 2 J pc can be calculated then an estimate of
the actual temperature distribution throughout the deforming material in
the vicinity of the die exit can be obtained. However, if the pattern of the
relative temperature distribution is sufficient then the material constant, {),
can be omitted from the calculations and then each of the temperature jumps
is proportional to the corresponding quotient ( V*/vN).It is then convenient
to equate the maximum temperature to 100 and the relative temperature
in each of the other regions is increased in proportion.
Measurements obtained from both the physical diagram of figure 10.20(a)
drawn to a scale H = 10 em and the hodograph of figure 10.20(b) drawn to a
scale 8 em= unit velocity are presented in table 10.2. These tabulated values
permit an upper bound estimate for the dimensionless extrusion pressure to
be obtained and also the value of the quotient (V* /vN) to be calculated for
each line of tangential velocity discontinuity or thermal discontinuity shown
in the physical diagram of figure 10.20(a). The hodograph of figure 10.20(b)
indicates the construction required for the determination of the value of vN
for each line of thermal discontinuity. From table 10.2 the value ofl:sV* =
1.534H. The dimensionless extrusion pressure is given by
Pef2k = IsV*/2H ~ 1.534H/2H ~ 0.767 (10.54)

Table 10.2 Measurements from the physical diagram offigure 10.20 (a) and
the hodograph offigure 10.20(b)
Line of velocity
discontinuity
or
thermal Tangential velocity
discontinuity Length,s discontinuity, Jl'l' sJI'I'

0.1H ab = 5.8/8 = 0.725 7.7/8 = 0.963 0.753 0.508H


0.31H ~: 5.8/8 : 0.725 6.9/8 = 0.863 0.840 0.268H
0.1H Q£1 - 1.5/8 - 0.188 8.7/8 = 1.09 0.172 0.132H
0.1H cd = 7.15/8 = 0.894 11.25/8 = 1.41 0.634 0.626H

l:sV* 1.534H

Table 10.3 Relative temperatures-see figure 10.20(a)


Region B

0.753 0.925 1.559 0.840 1.474


46 59 100 54 95
390 ENGINEERING PLASTICITY

Relative temperatures in the various regions and the values corrected on


the basis ofthe highest temperature being equated to 100 are given in table
10.3. The approximate pattern of relative temperature distribution is that
shown in figure 10.20(a).
Tanner and Johnson 13 have determined the pattern of relative temperature
distribution for plane strain extrusion through a smooth square face die for
a fractional reduction R = 2/3. Bishop 14, using a relaxation technique, has
investigated one particular fractional reduction for the plane strain extrusion
through a square face die allowing for thermal conduction.

10.10 PLASTIC BENDING OF A NOTCHED BAR-THE PLASTIC


HINGE
An interesting application of the limit theorems is the determination of
upper and lower bounds for the pure bending moment required to cause
plastic bending of a single notched bar. This problem also introduces the
concept of a plastic hinge.
Rotating rigid regions

Shearing occurs along Stationary rigid region


these circular arcs 'plastic hinge'

(a)

Stress free Stress free

Lines of stress discontinuity

(b)

Figure 10.21. Plane strain bending of a notched bar: (a) upper bound estimate
for the bending moment, M, assuming the existence of a plastic hinge; (b) a lower bound
estimate for the bending moment, M, involving lines of stress discontinuity
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 391
The bar shown in figure 10.21(a), which has a single notch, is subjected
to a pure bending mo11,1ent, M, such that incipient yielding occurs. The
minimum thickness of the bar is a and it is assumed that the width of the bar
is large so that the mode of deformation is plane strain.
An upper bound estimate for the pure bending moment, M, may be
obtained by one possible solution where it is assumed that yielding occurs
by shearing along the two circular arcs, ABC and ADC, which have a radius
of curvature, R, and subtend an angle, 2ex, at their centres. The rigid regions
of the bar, I and II, rotate about the stationary, rigid region ABCDA with
an angular velocity, w, in such a manner that the material shears along the
circular arcs. This system constitutes a plastic hinge for which there is now
considerable experimental evidence.
The magnitudes of the tangential velocity discontinuities which occur
across the circular arcs are V18 c = V1oc = Rw and the length of each circular
arcs= 2Rex. The rate of internal energy dissipation
dE/dt = ~)sV* = 2ksV* = 2k.2Rex.Rw = 4kR 2 cxw
and the rate of external work done, d »Jdt, by the equal and opposite applied
bending moments, M = 2Mw. Therefore
2Mw = 4kR 2 exw
or M =2kR 2 ex
However R = a/2sin ex
Hence M = ka 2 ex/2 sin 2 ex (10.55)
dM/dex = (ka 2 /2){ (sin 2 ex- 2ex sin ex cos ex)/sin 4 ex} (10.56)
Ifthe bending moment, M, is to have a minimum value then
sin 2 ex- 2ex sin ex cos ex= 0
or tan ex= 2ex
and
The best upper bound estimate for the bending moment is then given by
M =0.69ka 2 (10.57)
A lower bound to the bending moment is obtained from the first extremum
principle. The material on each side of the notch in the bar shown in figure
10.21(b) is stress free and the material, of thickness a, below the notch is
subjected to uniplanar bending. In this case, the neutral surface is half-way
between the root of the notch and the lower surface of the bar. When the
bar yields, the stress due to plastic bending is - 2k (compressive) above the
neutral surface and + 2k (tensile) below the neutral surface.
Stress discontinuities were discussed in section 7.3.7, where it was shown
that the component of stress normal to a line of stress discontinuity and also
392 ENGINEERING PLASTICITY

the shear stress tangential to the line must be continuous so that stress
equilibrium is maintained. However, a discontinuity in normal stress can
occur in a direction parallel to a line of stress discontinuity. The dash line and
the long chain line representing the neutral surface in figure 10.21(b) are
such lines of stress discontinuity. Consequently, the normal component of
stress parallel to these lines is subjected to a jump as the lines are traversed.
Thus, on progressing from the upper surface of the stressed bar to the neutral
surface the jump is from 0 to - 2k and in progressing from the neutral surface
to the lower surface of the bar the jump is from - 2k to + 2k. The internal
resisting moment, Mr, is Mr = 2k(af2)(af2) = 0.5ka 2 • For equilibrium with
regard to rotation, the externally applied bending moment, M, must be-
equal in magnitude but opposite in effect to the internal resisting moment.
Therefore
M=0.5ka 2 (10.58)
The bounds for the bending moment can therefore be stated as
0.5 ~ M/ka 2 ~ 0.69 (10.59)
For comparison, the slip line field solution for this problem presented by
Green 15 gives
M/ka 2 =0.63 (10.60)

10.11 UPPER BOUND ESTIMATES INVOLVING UNIT


DEFORMING REGIONS (KUDO UPPER
BOUND ESTIMATES)
Upper bound solutions to some plane strain metal forming problems, based
on the same principle as that discussed in section 10.7 but using a different
approach, have been obtained by Kudo 16 who introduced the concept of a
unit deforming region. For plane strain deformation he suggests a unit
rectangular deforming region and three main types of kinematically admissible
velocity field in the deforming regions are considered. These are (a) parallel,
(b) centre fan and (c) rigid triangle and it appears that the rigid triangle
velocity field, in which the deformation regions are bounded by straight lines,
gives the best solution for plane strain deformation. This is because the rate
of internal energy dissipation for this field is less than that for the other two.
A simple example of the use of the rigid triangle velocity field to obtain
a Kudo upper bound to the mean platen pressure during plane strain
compression is illustrated in figure 10.22. The physical diagram of figure
10.22(a) shows a rectangular section metal block of unit width and thickness
a, which is subjected to plane strain compression but is constrained by a
rigid, vertical boundary at its left-hand vertical face and the lower platen is
stationary. The velocity field consists of three rigid triangles which are
designated A, B and C.
It is assumed that the block deforms in such a manner that the rigid
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 393
Unit velocity

(a)

(b)

Figure 10.22. (a) Physical diagram showing a unit rectangular deforming


region consisting of a rigid triangle velocity field; (b) the corresponding hodograph

regions slide relative to each other and the top surface of the block 123
descends vertically under the influence of the moving upper platen with unit
velocity. The corresponding hodograph is given in figure 10.22(b). For
constancy of volume 1.1 =aU or U = 1/a, where U is the horizontal compo-
nent ofthe velocity of the region C and 1/a is the width to thickness ratio of
the rectangular section block.
The slip resistance per unit area of boundary between the deforming region
and the surrounding body such as tooling or per unit surface area of tangen-
tial velocity discontinuity is denoted by F = fY, wherefis referred to as the
slip resistance ratio and Y is the plane strain yield stress of the deforming
material. The value of the slip resistance ratio for a perfectly smooth interface
between the deforming material and tooling is zero and for a perfectly rough
surface or at a surface of velocity discontinuity f = 1/2 or 1/3 112 depending
on whether the Tresca or von Mises yield criterion is assumed to be appli-
cable.
If the material is rigid-perfectly plastic, the rate of total internal energy
dissipation is obtained by the summation of the rates of internal energy
dissipation at the assumed velocity discontinuities and at perfectly rough
394 ENGINEERING PLASTICITY

boundaries. Therefore
E= dE/dt == .Y[f15(a.1) + f 25 (afsin 0)(1/sin 0) + f 2iafsin </>)(1/sin </>)
+! 34(a.1) +! 54 .1(/3/a) +! 23 (1- /3)(1/a)]
= Y[f15 a + f 25 (afsin 2 0) +f 24(a/sin 2 </>) +! 34a +! 54(/3/a)
+ ! 23 {(1 - /3)/a}] (10.61)
It should be noted that f 15 ,f54 ,f23 and f 34 will have zero values if the
interfaces at these boundaries are perfectly smooth or 1/Y 12 if the interfaces
are perfectly rough and assuming von Mises yield criterion to be applicable.
Kudo defmes the coefficient of internal energy dissipation, e, as the rate of
total internal energy dissipation, E, divided by Y for a unit rectangular
deforming region which for the rigid triangle velocity field of figure 10.22(a)
is given by
e= [j15 a + f2s(ajsin 2 0) + f2iafsin 2</>) +f34a + fs4(f3/a) +f23{ (1- /3)/a}]
(10.62)
From the physical diagram of figure 10.22(a) :
sin 2 0 = a2f(a 2 + /3 2) and sin 2 4> = a2f{a 2 + (1- {3) 2}
Therefore
e= a{f1s +f34 + (2/3 112)} + (1/a){f23(1- /3) + fs4f3
+ (1/3 112)(1 - 2{3 + 2/32)} (10.63)
since f 24 = f 25 = 1/3 112.
Equation (10.63) can be differentiated with respect to the length, {3, and the
value of f3 determined to minimise the coefficient of internal energy dissipa-
tion, e, to yield the best upper bound solution. In general
E: = dE/dt = Yie;A; v; (10.64)
v;
where A; and are the projected area and velocity of the boundary, respec-
tively, of the ith unit deforming region and i is the number of unit regions
considered. The mean platen pressure is then given by
(10.65)
where A is the projected area on a plane normal to the direction of motion of
the tooling and Yo is the velocity of the tooling.
Kudo 16 has considered upper bound estimates for plane strain forging,
extrusion, extrusion forging and heading using suitable kinematically
admissible velocity fields involving more complex patterns of rigid triangles.
The concept of the unit rectangular deforming region for plane strain
deformation is extended by Kudo 17 to axisymmetric forming problems by
introducing the unit cylindrical deforming region. This upper bound approach
has proved useful because the exact mathematical analysis of axisymmetric
plastic deformation is possible only by making certain additional assump-
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 395
Unit velocity

3 t

Figure 10.23. A unit cylindrical deforming region consisting of a parallel velocity


field (after Kudo 17)

tions concerning the stress state even for a rigid-perfectly plastic material.
Also the analysis of complicated problems is simplified to some extent.
In figure 10.23, the unit cylindrical deforming region has a rectangular
section, the sides of which are parallel and normal to the axis of symmetry,
respectively. This region has an external radius of unity, an internal radius,
b, and a height to external radius ratio of a. It is assumed that the mode of
deformation is such that its top surface 13 descends with unit velocity and, for
this particular case, the inner surface 34 moves inwards as a straight vertical
line. The surfaces 12 and 24 are stationary.
Two types of kinematically admissible velocity field may be considered.
In the parallel velocity field, the strain rate of an element is independent
of its vertical coordinate. The other field is the triangular velocity field in
which the unit deforming region is assumed to consist of two or three annular
parts having triangular sections. Velocity fields of this type are given in
figure 10.24. All the sides of these triangular sections are lines of velocity
discontinuity and are assumed to move as strailght lines without changing
their inclinations. The radial component of velocity in each section is
assumed to be independent of z.
The rate of internal energy dissipation in these two types of velocity field
396 ENGINEERING PLASTICITY

(a)

(b)

(c)

Figure 10.24. Rigid triangle velocity fields for a cylindrical deforming region:
(a) compression of an annulus with inward radial flow; (b) compression of an annulus
with outward radial flow; (c) compression of a solid circular cylinder (after Kudo 17)

is given by

.
E = dE/dt = (2/3) 112 Y vf {e; + ei + e; + (1/2)Y;z}IndV + Y f/sds (10.66)
where the first integration is performed throughout the entire volume of the
deforming region and the second integration is performed over all surfaces of
velocity discontinuity inside the material and at material-tool interfaces.
The rate of relative slip on a surface of velocity discontinuity is denoted by S.
Considering a parallel velocity field in the unit cylindrical deforming
region shown in figure 10.23 then for constancy of volume n(1 - r 2 ).1 =
- u.2nra and - wn(1- r 2 ) = - u.2nrz. The velocity components of an
element at radius r are thus
u= - (1 - r 2 )/2ar, w= - zja (10.67)
and the strain-rate components are given by
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 397
er = fJujfJr = (1 + r 2 )/2ar2 )
66 = (1/r){(fJvjfJO) + u} = u/r = - (1- r 2 )/2ar2 (10.68)
ez = -1/a
Yrz = 0
Using the values of the strain-rate components of equations (10.68)

I {e;
v + ei + e; + (Y;z/2)} 112 d v = (1/2 i a)
1 2 I v { (1 + 3r4 )/r4 pi 2 d v
An elemental volume, d V = rdO. dr. dz. Therefore

I
(1/2 1i 2 a) v {(1 + 3r4 )/r4 } 1 i 2 dV = (1/2 1i 2 a) J:'J: I {(1 + 3r4 )/r4 } 1 i 2

x rdO.dr.dz

=2 112 n I{(1+3r4 ) 112/r}dr

= (2 112 n/4)[4- 2ln 3 + 2ln ( {(1 + 3b4 ) 1 i 2


+ 1} /b 2 ) - 2(1 + 3b4 ) 1i 2 ]

Also I /sdS = I(/13 + / 24){(1- r 2 )/2ar}2nrdr +I: (/12 + bf34)(z/a)2ndz

= (n/a)(f13 + / 24) [ (1 -b)- {(1 - b3 )/3}] + na(f12 + bf34 )


Therefore
E = (nY/2.3 112 )[4- 2ln 3 + 2ln ({ (1 + 3b4 ) 1 i 2 + 1}/b2 ) - 2(1 + 3b4 ) 1i 2 J
+ (Yn/a) [ (/13 + / 24)(1 -b)- (f13 + f 24){ (1 - b3 )/3}
+ a 2 ( /12 + b/34)] (10.69)
The coefficient of internal energy dissipation
(10.70)
where the denominator is the product of the uniaxial yield stress, Y, the
surface area which is compressed, 1t(1 - b2 ), and the compression velocity
which is equal to unity. It should be noted that this quantity remains
unchanged if the same annulus is compressed radially on its cylindrical
surface when its top surface 13 will ascend in contact with the rigid platen
if this moves freely in the axial direction. Hence
e= {1/3 112(1- b2 )} [2 -In 3- (1 + 3b4 ) 1i 2 +In ({ (1 + 3b4 ) 1i 2 + 1} /b 2 ) J
+ {(f13 + / 24)/3a(1 +b) }(2- b- b2 ) + {a/(1 - b2 ) }(/12 + b/34)
(10.71)
which can be stated as
e= M(b) + (/13 + / 24)N(b)(1/a) + (f12 + bf34)a/(1- b2 ) (10.72)
398 ENGINEERING PLASTICITY

where M(b) = {1/3 112(1- b2 )} [2 -In 3- (1 + 3b4 ) 1 i 2


+In <{(1 + 3b4 ) 1 i 2 + 1}jb2 ) J
(10.73)
and N(b) = (2- b- b2 )/3(1 +b) (10.74)
The value of M(b) approaches infinity as b decreases and approaches 2/3 112
as b approaches unity.

10.12 UPPER BOUND ESTIMATES FOR AXISYMMETRIC


DEFORMATION (KOBAYASHI UPPER BOUND
ESTIMATES)
In section 10.11 the analysis is given for a Kudo upper bound solution to an
axisymmetric compression problem assuming a parallel velocity field.
Solutions are also given by Kudo 1 7 for axisymmetric forming problems
assuming triangle velocity fields where conical surfaces are introduced as
surfaces of velocity discontinuity in the cylindrical deforming region corres-
ponding to straight lines on a meridian plane. However, these triangle fields
were found to give better upper bound solutions only over a limited range of
conditions when compared with solutions obtained assuming a parallel
velocity field. This situation arises because, unlike for plane strain deforma-

(a)

-
(b)

Figure 10.25. Velocity fields for a cylindrical deforming region with curved
lines of velocity discontinuity on a meridian plane: (a) compression of an annulus with
inward radial flow; (b) compression of an annulus with outward radial flow (after
Kobayashi 18 )
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 399
Unit velocity

•I

Figure 10.26. Compression of an annulus within a rough container by a rough


platen. The assumed velocity field consists of curved lines of velocity discontinuity on a
meridian plane (after Kobayashi 18)

tion, the deforming material cannot be described completely by assuming


rigid body motion. Therefore, in the Kudo analyses only simple patterns of
velocity discontinuity are considered.
With a view to overcoming this limitation, Kobayashi 18 • 19 proposed
improved kinematically admissible velocity fields in which the lines of
velocity discontinuity on a meridian plane are assumed to be curved.
Examples of these patterns of velocity discontinuity for radially inward and
outward flows in a cylindrical deforming region are shown in figure 10.25.
In the velocity fields of figure 10.25(a) for inward flow it is assumed that the
axial component of velocity is constant in each region and for the outward
flow of figure 10.25(b), the radial component of velocity is assumed to have
the form u = Arm.
To demonstrate the method of determining the velocity components and
the total rate of internal energy dissipation, compression of an annulus
within a perfectly rough container by a perfectly rough platen, as shown in
figure 10.26, will be considered. The deforming material is assumed to be
rigid-perfectly plastic and the von Mises yield criterion and associated flow
rule are applicable.
A simple pattern of curved surfaces of velocity discontinuity is assumed
in the cylindrical deforming region consisting of three regions in which
the radial and axial velocity components are
Region 1 :
Region 2: (10.75)
Region 3 : u3 = - (1 - b2 )/2ar, v3 = -a
The velocity field defined by equations (10. 75) satisfies the incompressibility
equation
400 ENGINEERING PLASTICITY

(oujor) + (ujr) + (ovjoz) = 0


and also the velocity boundary conditions.
The equations defining the curves of velocity discontinuity can be deter-
mined since the component of velocity normal to these curves must be
continuous across the curves. For example, the velocity components along
the curve of velocity discontinuity 45 must satisfy the equation
(u 3 - u2 )tan e= v3 - V2

e
where tan = dz45 /dr. Therefore
dz 45 /dr = 2araj(1 - b2 )

and z45 = J: {2ara/(1 - b2 )} dr = aa { (r 2 - b 2 )/(1 - b2 )} (10.76)

Similary z 35 = {a(1- a)(1- r 2 )/(1- b2 )} + aa (10.77)


The rate of internal energy dissipation for the continuously deforming
region 345 is given by

£345 = ff ai2nrdz = (na/3 112 ){2ln(1/b)- (1- b 2 )} (10.78)

where a is the equivalent stress and i is the equivalent strain rate.


The rate of internal energy dissipation along the curved velocity dis-

f
continuity 35 is

E35 = (2na /3 112 ) r[V3*5 ] dS

= (na /3 112 ) [ (4/3){ (1 + b + b 2 )/(1 +b) }(1 - a) 2 a+ (1 - b2 )(1 -- b)(1/a)]


(10.79)
where V3*5 is the velocity discontinuity along the curve 35 and Sis the distance
along the curve. Similarly
E45 = (n:0:/3 112 ) [ (4/3){ (1 + b + b 2 )/(1 +b)} a 2 a+ (1 - b2 )(1 - b)(1/a) J
(10.80)
and along the rough surface 15
E15 = (2n0'/3 112 )(1 - a)a (10.81)
If the total rate of internal energy dissipation, obtained by adding equations
(10.78) to (10.81), is divided by an(1 - b2 ).1 the mean platen pressure, p, is
then given by
p/0'=(1/3 1i 2 )[{2/(1-b 2 )}ln(1/b)-1]
+ {2a/(1 - b2 ) }[ (2/3.3 112 ) { (1 + b + b2 )/(1 +b) }(1 - 2a + 2a 2 )
+ (1 - a)/3 1 12 ] + 2(1 - b)j3 112 a (10.82)
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 401
Equation (10.82) can be differentiated with respect to cx and the value of
cx determined to minimise the dimensionless mean platen pressure, p;a, and
yield the best upper bound solution. This value of cx is found to be
0( = (1/2)[1 + (3/4){ (1 + b)/(1 + b + b2)}] (10.83)
Other patterns of velocity field involving curved surfaces of velocity
discontinuity have also been considered by Kobayashi for the compression of
cylindrical deforming regions. The method has also been applied to bar
extrusion and wire drawing. For some axisymmetric deformation problems,
the velocity fields proposed by Kobayashi produce better upper bound
solutions for the forming parameter than the rigid triangle velocity fields
suggested by Kudo.
The methods of analysis described by Kudo and Kobayashi and outlined
in sections 10.11 and 10.12 to obtain an upper bound estimate ofthe required
forming parameter have been applied by McDermott and Bramley 20 to the
press forging of complex shapes having axial symmetry. A forging is consi-
dered to be divided into a combination of elemental rings so that the
procedure extends the basic approach to include regions which have rectan-
gular, triangular, convex circular and concave circular sections. The necessary
calculations are programmed into a computer to describe the flow stress,
frictional conditions, geometry, etc., for each deforming region to obtain an
output of the total forging force required. The programme also enables
the optimum flash geometry to be predicted.

10.13 UPPER BOUND ESTIMATES FOR THE PLASTIC


BENDING OF TRANSVERSELY LOADED THIN
PLATES

The method employed in the previous sections to obtain an upper bound


to the forming parameter required to produce plastic flow derived from the
second extremum principle, is now extended to analyse thin flat plates which
are transversely loaded and supported in some manner. The concept of the
plastic hinge, discussed in section 10.10, together with an acceptable collapse
mechanism are utilised to account for incipient plastic yield. The method
discussed in this section was originally presented by Johnson 21 and has
been rigorously justified by Collins 22 •
A kinematically admissible velocity field is proposed which defines the
angular velocity of rotation for each element of the plate and a diagram of
angular velocities or hodograph can then be constructed. The rate of internal
energy dissipation due to the fully plastic bending moments occurring in the
plate and at its boundary is then determined and the rate at which external
work is done by the loading is equated to the rate of internal energy dissi-
pation to obtain an upper bound estimate for the required parameter.
402 ENGINEERING PLASTICITY

It is assumed that the plate is rigid-perfectly plastic. Deflections due to


elastic deformation are therefore absent or may be considered insignificant
compared with deflections due to plastic deformation. However, since
incipient yield is assumed, difficulties concerned with deflection and the
change in configuration of the plate need not arise. Only the effects of
bending moments are considered and therefore membrane stress and shear
stress distributions are neglected. Since engineering metals are generally
strain-hardening at ambient temperature the estimated parameter is not
likely to be a gross overestimate.

10.13.1 An annular plate perfectly clamped along its outer edge and
uniformly transversely loaded along its inner edge

Consider a thin annular plate of uniform thickness, t, having an external


diameter, 2a, and internal diameter, 2b, to be perfectly clamped along its
outer edge and subjected to a uniformly distributed transverse load ofF per
unit length along its inner edge as shown in figure 10.27(a).
The annular plate may be considered to consist of a very large number
of identical elemental sectors, designated 1, 2, 3, 4 ... , where each sector sub-
tends a small angle, [)(), at the centre of the plate. It is assumed that when
incipient yield occurs due to the loading of the plate, a plastic hinge is
developed at the external boundary of each sector, where it is clamped, and
additional plastic hinges are developed at the interfaces between the separate
sectors where they are joined along their radii. If the internal circumference
of the plate is designated A, B, C, D, E... and the external circumference by
a, {J, y, f>, e... then sector 1 is defined as AB{Ja and a plastic hinge exists at afJ
and also along the interface BfJ between sector 1 and sector 2, etc.
When plastic collapse occurs let the inner circumference of the plate
descend with a linear velocity, u, as each elemental sector 1, 2, 3, 4 ... rotates
as a rigid body about the plastic hinges afJ, {Jy, yf>, f>e .•• , respectively, with
an angular velocity ro = u/(a- b), as indicated in figure 10.27(a). The angular
velocity, ro, of each sector can be represented by a vector of magnitude,
u/(a - b), drawn to a suitable scale, parallel to the appropriate axis of
rotation and the sense of the vector defined in accordance with, say, the
right-hand screw rule. For example, sector 1 rotates anticlockwise about the
plastic hinge, r:x{J, at its external boundary when viewed from a towards fJ
with an angular velocity ro = u/(a- b). The vector representing this angular
velocity is shown in figure 10.27(c) as 01 which is drawn to a suitable scale to
represent the magnitude, ro, and is parallel to the axis of rotation, rx{J, with
a sense which is from 0 to 1. In a similar manner, the angular velocity vectors
for the other sectors 2, 3, 4, ... are represented by 02, 03,04 ... in figure 10.27(c).
The diagram thus constructed is referred to as a diagram of angular velocities
or simply as the hodograph. It should be noted that in the hodograph of
figure 10.27(c), the vectors 12, 23, 34 ... represent the rate of relative rotation
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 403

·~

(a)

(b)

'" (if) 1234

Figure 10.27. (a) Annular plate perfectly clamped at its outer edge subjected to a
uniformly distributed transverse load at its inner edge; (b) annular plate divided into a
very large number of elemental sectors which subtend an angle MJ at the centre of the
plate; (c) corresponding diagram of angular velocities or hodograph to physical dia-
gram (b); (d) the hodograph becomes a circle when f>(}-+ 0

between one sector and the adjacent sector. For example, the angular
velocity vector, 01, can be resolved into two rectangular components, On and
nl. Similarly, the angular velocity vector, 02, can be resolved into the two
components, On and n2. The component, On, is the same for the sectors 1
and 2. It follows that they both rotate with the same angular velocity about
an axis perpendicular to the radial plastic hinge, B/3, at their interface.
The total rate of change of angle between sector 1 and sector 2 across the
radial hinge, B/3, is given as the vector difference of nl and n2, that is, a
vector of magnitude 1121 which represents an angular velocity discontinuity
across the radial hinge B/3. As the angle (j(} approaches zero, each sector
becomes infinitely small and, in the limit, the hodograph becomes a circle of
radius, w, as shown in figure 10.27(d).
404 ENGINEERING PLASTICITY
The rate of total internal energy dissipation is therefore given as the
sum of (a) the rate of internal energy dissipation at all the plastic hinges
along the perfectly clamped external circumference of the plate and (b) the
rate of internal energy dissipation at all the radial hinges. Hence
dE/dt = '2:.a.f3.w.Mp + '2:.Bf3.1121.Mp (10.84)
where M P is the fully plastic bending moment per unit length. Therefore
dE/dt = 2nawMP +(a- b)2nwMP = 2nwMP(2a- b) (10.85)
The rate of external work done by the uniformly distributed load F per unit
length is
dWjdt = F.2nb.u (10.86)
Therefore F.2nb.u = 2nawMP(2a- b)
or F = (2a- b)MP/b(a- b) (10.87)
The total line force, F*, to cause plastic collapse is
F* = 2nbF = 2n(2a- b)MP/(a- b) (10.88)
If the plate is solid such that b = 0 then the force concentrated at the
plate centre required to cause plastic collapse is
F* =4nMP (10.89)

10.13.2 An annular plate perfectly clamped along-Its outer edge


subjected to a uniform lateral pressure
Assuming that the same conical mode of deformation, first referred to in
section 9.5.2, is applicable and the annulus of figure 10.27(a) to be subjected
to a uniform lateral pressure, p, then an upper bound to the limiting lateral
pressure, p*, may be determined by following the same method used in
section 10.13.1.
An element of the plate at any radius, b : : :; r : : :; a, descends with a linear
velocity
ur = u(a- r)j(a- b) (10.90)
The rate of external work done by the limiting lateral pressure, p*, is

d wjdt = f: p*. 2nrdr. ur = {2np*uj(a- b)} J: (a- r)rdr

= (np*uj3)(a- b)(a + 2b) (10.91)


The rate of internal energy dissipation at the circumferential and radial
plastic hinges is given by equation (10.85). Therefore
(np*uj3)(a- b)(a + 2h) = 2nwMP(2a- b) (10.92)
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 405
and the limiting lateral pressure, p*, is given by
p* = 6MP(2a- b)/ {(a- b) 2 (a + 2b)} (10.93)
For a solid circular plate with b = 0
p* = 12MP/a2 (10.94)
which may be compared with the value deduced in section 9.5.2(b) and
given by equation (9.66).

10.13.3 An annular plate position fixed at its outer edge subjected


to a uniform lateral pressure
The situation to be analysed is that illustrated in figure 10.28 and the approach
is similar to that adopted in the previous section. However, the expression
for the rate of internal energy dissipation given by equation (10.85) must be
modified because plastic hinges are not developed at the outer edge of the
plate since the plate is free at this boundary and the fixing moment is zero.
Equation (10.85) therefore reduces to
dE/dt =(a- b)2nwMP (10.95)
and then
(np*u/3)(a- b)(a + 2b) =(a- b)2nwMP
or p* = 6Mp/{(a- b)(a + 2b)} (10.96)
For a solid circular plate simply supported at its outer edge with b = 0
p* = 6MP/a 2 (10.97)
which is identical with equation (9.60).

Figure 10.28. Annular plate position fixed at its outer edge subjected to a uniform
lateral pressure

10.13.4 A rectangular plate position fixed at its boundary subjected to a


uniform lateral pressure
Consider a thin rectangular plate of uniform thickness having side lengths
of 2a and 2b, where b ~ a to be position fixed over its entire boundary
406 ENGINEERING PLASTICITY

ABCD and subjected to a uniform lateral pressure, p, as indicated in figure


10.29.
An upper bound estimate of the limiting lateral pressure, p*, depends
on the mode of deformation assumed and hence the proposed angular

Figure 10.29. Rectangular plate position fixed at its entire boundary subjected
to a uniform lateral pressure

Hodograph
Ia I

(b)

Hadagraph

Figure 10.30. Different modes of deformation for a rectangular plate: (a) plastic
hinges coinciding with plate diagonals; (b) conical mode of deformation; (c) mode of
deformation involving five plastic hinges
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 407
velocity field. There are a number of different possibilities of which three
valid modes will be considered. These are shown in figure 10.30. In figure
10.30(a) it is assumed that the plastic hinges coincide with the plate diagonals
AC and BD. A conical mode of deformation is illustrated in figure 10.30(b)
and in figure 10.30(c) it is assumed that five plastic hinges are developed
along AE, DE, BF, CF and EF.
For case (a), when plastic collapse occurs, the point P which is the inter-
section of the plate diagonals in figure 10.30(a) descends with a velocity,
u, and therefore the centre of gravity of each triangular section, such as ABP,
descends with a velocity of u/3. The length of each plastic hinge, such as AP,
is (a 2 + b2 ) 1' 2 and the angular velocity discontinuity across each plastic
hinge is given by w* in the hodograph where w* = (wi + wD 112 •
The rate of total internal energy dissipation occurring at the four identical
plastic hinges is therefore given by
dEjdt = 4MP(a 2 + b2 ) 112 (wi + w~) 1 1 2
The rate of external work done by the uniform limiting lateral pressure,
p*, is
dWjdt = p* .2a.2b.uj3
Therefore p* .2a.2b.uj3 = 4MP(a 2 + b2 ) 112 (wi + w~) 1 1 2
However u= aw 1 = bw 2
Hence p* = 3MP(a 2 + b2 )/a 2 b2 (10.98)
which can be rewritten as
p* = (6MP/a 2 )[ { 1 + (a 2 jb 2 )} /2] (10.99)
For a square plate such that a= b
p* = 6MP/a 2 (10.100)
For case (b), shown in figure 10.30(b), the similarity with that considered
in section 10.13.3 will be noted. As given by equation (10.97), p* = 6MP/a 2 •
For case (c), shown in figure 10.30(c), the rate of total internal energy dissi-
pation at the plastic hinges is
dEjdt = [(4.2 112 a.2 112 w) + {2(b- a).2w} ]MP
=4w(a + b)MP
The rate of external work done by the uniform limiting lateral pressure,
p*, is
dW jdt = p*[ {4.2a(a/2).(u/3)} + {2.2(b- a)a.(u/2)} J
= p*[(4a 2 uj3) + 2a(b- a)u]
= p*(4a 2 uj3) [ 1 + {3(b- a)j2a}]
Therefore p*(4a 2 uj3) [ 1 + {3(b- a)j2a} J= 4w(a + b)MP
408 ENGINEERING PLASTICITY

However u=aw
Therefore p*(4a 3 /3)[1 + {3(b- a)/2a}] = 4(a + b)MP
or p* = (6MP/a 2 ){ (a+ b)/(3b- a)}
= (6MP/a 2 ) [ { 1 +(a/b) }/{3- (a/b)} J (10.101)
If the plate is square such that a= b then equation (10.101) again reduces
top*= 6MP/a 2 •
It is found that the mode of deformation assumed for case (c) yields a
lower value for the upper bound estimate of the limiting lateral pressure, p*,
for values 0 ~ ajb ~ 1, given by equation (10.101), than either of the other
two modes where p* is given by equations (10.99) and (10.97), respectively.
If instead of assuming angles such as ADE are 45° they are designated fJ,
then it can be shown that
p* = (6M P/ a 2 ) [ {1 +(a/b) cot fJ} I {3 -(a/b) tan fJ}] (10.102)
and p* has least value when
tan fJ = {3 + (a/bf} 112 - (a/b) (10.103)
Combining equations (10.102) and (10.103) yields
p* = (6MP/a 2 )[ {3 + (a/bf} 1 i 2 - (a/b)]- 2 (10.104)
If a/b approaches zero, that is, a very long strip then both equations (10.101)
and (10.104) reduce to
(10.105)
Other shapes of plates with different loading and support arrangements
have been analysed by Johnson 21 •

REFERENCES
1. Gvozdev, A. A., Calculation of load bearing capacity of structures by
the method of limiting equilibrium (in Russian), Stroyisdat, U.S.S.R.
(1949)
2. Gvozdev, A. A. (translated by R. M. Haythornthwaite), The determina-
tion of the value of the collapse load for statically indeterminate systems
undergoing plastic deformation, Int. J. mech. Sci., 1, 322 (1960)
3. Hill, R., The Mathematical Theory of Plasticity, ch. III, Oxford Univer-
sity Press, London (1950)
4. Koiter, W. T., General theorems for elastic-plastic solids, in Progress in
Solid Mechanics (ed. by Sneddon and Hill), vol. 1, North-Holland,
Amsterdam (1960)
5. Drucker, D. C., Greenberg, W. and Prager, W., The safety factor of an
elastic-plastic body in plane strain, Trans. Am Soc. mech. Engrs, 73,
J. appl. Mech., 371 (1957)
EXTREMUM PRINCIPLES FOR A RIGID-PERFECTLY PLASTIC MATERIAL 409
6. Koiter, W. T., Stress-strain relations, uniqueness and variational
theorems for elastic-plastic materials with a singular yield surface,
Q. appl. Math., 11, 350 (1953)
7. Johnson, W., Estimation of upper bound loads for extrusion and
coining operations, Proc. Instn mech. Engrs, 173, 61 (1959)
8. Slater, R. A. C., Velocity and thermal discontinuities encountered
during the forging of steels, Proc. Manchester Association of Engineers,
No.5 (1965-66)
9. Green, A. P., The compression of a ductile material between smooth
dies, Phil. Mag., 42, 900 (1951)
10. Johnson, W., de Malherbe, M. C. and Venter, R., Upper bounds to the
loads for the plane strain working of anisotropic metals, J. Mech.
Engng Sci., 14,280 (1972)
11. Massey, H. F., The flow of metal during forging, Proc. Manchester
Association of Engineers (November 1921)
12. Johnson, W., Baraya, G. L. and Slater, R. A. C., On heat lines or lines
of thermal discontinuity, Int. J. mech. Sci., 6, 409 (1964)
13. Tanner, R.I. and Johnson, W., Temperature distributions in some fast
metal-working operations, Int. J. mech. Sci., 1, 28 (1960)
14. Bishop, J. F. W., An approximate method for determining the tempera-
tures reached in steady motion problems of plane plastic strain, Q. Jl
Mech. appl. Math., 9, 236 (1956)
15. Green, A. P., The plastic yielding of notched bars due to bending,
Q. Jl Mech. appl. Math., 6, 223 (1953)
16. Kudo, H., An upper bound approach to plane strain forging and
extrusion-I, Int. J. mech. Sci., 1, 57 (1960)
17. Kudo, H., Some analytical and experimental studies of axisymmetric
cold forging and extrusion-I, Int. J. mech. Sci., 2, 102 (1960)
18. Kobayashi, S., Upper bound solutions of axisymmetric forming
problems-I, Trans. Am. Soc. mech. Engrs, J. Engng Ind., 122 (May 1964)
19. Kobayashi, S., Upper bound solutions of axisymmetric forming
problems-II, Trans. Am Soc. mech. Engrs, J. Engng Ind, 326 (November
1964)
20. McDermott, R. P. and Bramley, A. N., An elemental upper bound
technique for general use in forging analysis, Proc. 15th Intern. Mach.
Tool Des. and Res. Conf., p. 437, University of Birmingham, 18-20
September, 1974, Macmillan Press, London (1975)
21. Johnson, W., Upper bounds to the load for the transverse bending of
flat, rigid-perfectly plastic plates, Int. J. mech. Sci., 11,913 (1969)
22. Collins, I. F., On the analogy between plane strain and plate bending
solutions in rigid perfect plasticity theory, Int. J. Solids Struct., 7, 1037
(1971)
Appendix 1 · The Rule of Sarrus
for the expansion of
third-order
determinants

Consider the third-order determinant

The three columns of the determinant are written down and the first
two columns are repeated as follows:
Positive terms

""'~
al bl ~/
cl al/ bl/

a2
~XX/
b2 c2 a2 b2
/XX~
a3 b3 c3 a3 b3
// /~""'""'
Negative terms
The products of the elements on the lines which are drawn from the
top to the bottom and run from left to right give the positive terms in the
expansion. The products of the elements on the lines drawn from bottom
to top and run from left to right give the negative terms. Therefore, the
expansion of the determinant is given by
~ = a 1 b2 c3 + b1 c2 a 3 + c 1 a2b3 - a3 b2c1 - b3 c2a 1 - c3 a2b 1 (ALl)
Now considering equation (2.24)
(ax- sn) !yx !zx

!xy (ay- sn) !zy =0


!xz !yz (az- sn)

and expanding by the Rule ofSarrus


APPENDIX 411
Positive terms

Therefore

~-~~-~~-~+~~~+~~~-~~-~~
- ryz rzy(ax- sn)- (az- sn)rxy ryx = 0 (A1.2)
which agrees with equation (2.25).
Appendix 2: The character istics
of partial differenti al
equations

The slip lines previously discussed in chapter 7 are the characteristics of


the hyperbolic partial differential equations which define the problem of
plane strain deformation of a rigid-perfectly plastic material. This appendix
is intended to give the reader an insight into the mathematical basis for
slip line fields.

FIRST-ORDER PARTIAL DIFFERENTIAL EQUATIONS


Initially consider the first-order partial differential equation of the type
A(oujox) + B(oujoy) = c (A2.1)
where A, B and C are functions of u, x and y but not of the partial derivatives
of u. This type of equation is known as quasi-linear. Let
oujox = p and oujoy = Q (A2.2)
Equation (A2.1) can then be rewritten as
AP+BQ=C (A2.3)
Given some curve C, in the xy plane, along which the function u is specified
the problem may then be considered in terms of whether a solution of u
in equation (A2.1) exists which satisfies the initial conditions. The solution,
of course, depends on the curve C. This two-dimensional equivalent of
the ordinary first-order equation where the initial value is defined at a
point is the so-called Cauchy problem.
If values of u, on the curve C, are given and P and Q could be determined
on the curve such that either equation (A2.1) or equation (A2.3) was satisfied
then the value of u, a small distance away from the curve C, could be deter-
mined by means of the Taylor series expansion in two variables. Let the
values of x and y on the curve C be designated x = x( C) and y = y( C), then
u(x, y) = u {x(C),y(C)} + {x- x(C) }(oujox)c + {y- y(C) }(oujoy)c + ...
(A2.4)
APPENDIX 413
It follows that the value of u could therefore be determined at some
other neighbouring curve. If this process is repeated, the value of u could
be determined over a region, although the limits of the region are unknown
at present. The Cauchy problem, in this case, may therefore be defmed in
terms of determining the values of P and Q on the curve C.
Let s be the arc length along the curve C, then if the derivatives P = ouj ox
and Q = oujoy actually exist
oujos = (oujox)(dxjds) + (oujoy)(dyjds) = P(dx/ds) + Q(dyjds) (A2.5)
The two equations (A2.3) and (A2.5) are thus available to solve for the two
unknowns, P and Q, from which
P = (Cdy- Bdu)j(Ady- Bdx)}
(A2.6)
and Q = (Adu- Cdx)j(Ady- Bdx)
Given A, B and C, the value of u on the curve C and the shape of the curve,
the values of P and Q can be obtained from equation (A2.6) and the value
of u in the neighbourhood of the curve C can then be obtained from equation
(A2.4). This cannot be true for any curve C, because if the curve is such that
the denominator in equations (A2.6) vanishes, that is, if
dy/dx =B/A (A2.7)
then there is no solution unless the numerators also vanish. A curve C,
having an equation which satisfies equation (A2.7), is referred to as a charac-
teristic curve. If the value of u is specified along such a curve it follows that
there will be no solution to the problem unless the value of u is such that
dujdy = C/B or dujdx = C/A (A2.8)
in which case the numerators in equations (A2.6) vanish. If u is specified on
the curve C, satisfying equation (A2.8), there will be an infinite number
of solutions because equations (A2.6) are then indeterminate. Equation (A2.8)
is known as the compatibility equation.
Summarising, if the value of u is specified along a curve C, then a solution
to equation (A2.1) exists only if the curve C is not a characteristic of the
differential equation. If the curve C is a characteristic, then either there is
no solution when equation (A2.8) is not satisfied or there is an infmite
number of solutions if equation (A2.8) is satisfied.

SECOND-ORDER PARTIAL DIFFERENTIAL EQUATIONS


Now consider the second-order, quasi-linear partial differential equation
(A2.9)
A linear partial differential equation is linear in the dependent variables
and their partial derivatives, that is, the coefficients are constants or functions
of the independent variables. A quasi-linear partial differential equation is
414 ENGINEERING PLASTICITY

linear in the highest order derivatives but the coefficients may be functions
of the independent variables and the dependent variables and their deri-
vatives up to the second highest orders occurring in the equation. Thus
equation (A2.9) is linear if A, B, C and D are constants or functions of x and
y, and quasi-linear if they are also functions of any one or more of u, ou/ ox
and fJujfJy. Let
82 u/8x 2 = R, 82 u/8x8y = S and 82 uj8y 2 = T (A2.10)
then equation (A2.9) becomes
AR+BS+CT=D (A2.11)
Assume that on some initial curve, Ci, in the xy plane, the values of
the partial derivatives P and Q are given, then the value of u and its normal
derivative can be determined. The value of u, P and Q along the curve Ci
is known as a strip of first-order and u, P and Q must satisfy the equation
dufds = (8u/8x)(dx/ds) + (8u/8y)(dyjds)
or du=Pdx+Qdy (A2.12)
called the strip condition.
The problem now is to determine whether a solution for u, of equation
(A2.9), exists which satisfies the initial conditions. If the second partial deri-
vatives R, S and T on the curve can be determined which satisfy equation
(A2.9) then u can be determined as well as the first partial derivatives, P and
Q, a small distance away from the curve by means of the Taylor series
expansion as before. On the curve
du = (ou/ox)dx + (oufoy)dy
d(fJujfJx) = (8 2 u/8x 2 )dx + (8 2 u/8x8y)dy
or dP=Rdx+Sdy (A2.13)
and d(fJujfJy) = (8 2 uj8x8y)dx + (8 2 uj8y 2 )dy
or dQ=Sdx+ Tdy (A2.14)
Hence,if A, B, C, D, dp and dQ are known on the curve, the three equations
(A2.11), (A2.13) and (A2.14) are available to solve for the three unknown
second partial derivatives R, S and T.
For example
dx dP 0
0 dQ dy
A D c
S= (A2.15)
dx dy 0
0 dx dy
A B c
APPENDIX 415
If the determinant in the denominator vanishes there is no solution unless
the numerator also vanishes, when there is an infinite number of solutions.
Assuming the denominator vanishes
dx dy 0
0 dx dy =0
A B C
that is A(dyjdx) 2 - B(dyjdx) + C = 0 (A2.16)
which is a quadratic equation indy/ dx. Therefore
dyjdx = {B ± (B 2 - 4AC) 1' 2 }/2A (A2.17)
Equation (A2.17) defmes two families of curves which are the characteristic
curves and if the data are known along one of these curves no solution exists
unless, as stated before, the numerator also vanishes. The compatibility
equation for this case is
AdPdy + CdQdx- Ddxdy = 0 (A2.18)
which must be satisfied along the characteristic curves.
The original second-order; quasi-linear partial differential equation (A2.9)
was replaced by equations (A2.17) and (A2.18) and, if the initial curve is not a
characteristic, a solution can be determined by obtaining the characteristics
from equation (A2.17) and then integrating along the characteristics using
equation (A2.18). If the equation is non-linear, that is, A, Band Care functions
of u, P and Q, then equations (A2.17) and (A2.18) require to be solved simul-
taneously as for the first-order equation A2.1).
If B 2 - 4AC > 0, the roots of equation (A2.17) are real and there are two
characteristic curves passing through every point in the xy plane. Equation
(A2.9) is then classified as hyperbolic.
If B 2 - 4AC = 0, there is only one characteristic curve passing through
every point in the xy plane. Equation (A2.9) is then classified as parabolic.
If B 2 - 4AC < 0, equation (A2.9) is elliptic and there are no real charac-
teristic curves.
In the theory of plasticity the most important class of partial differential
equation is the hyperbolic. For a problem involving a hyperbolic partial
differential equation such as equation (A2.9) with appropriate boundary
and initial conditions known the determination of the field of characteristics
is a solution to the problem.
Discontinuities of the second partial derivatives of u are possible across
characteristic curves but not across any other curves. The value of u cannot
be defmed arbitrarily on a curve which intersects a characteristic more
than once. In general, no solution exists for a hyperbolic partial differential
equation such as equation (A2.9) within a closed curve in the xy plane if
u is defmed at all points on the curve.
Author Index*

Alder, J. F., 149, 150, 151, 172 Campbell, J. D., 149, 173
Alexander, J. M., 4, 10,208,236,249,250,259, Christopherson, D. G., 186, 259, 276, 305
260,261,294,296,308,309,338,346 Chung,S. Y.,295,309,328,343,346,347
Aris, R., 61, 66 Coffin, L. F., 145, 172
Ashcroft, K., 307, 308 Cole, I. M., 295, 309
Atkins, A. G., 289,308 Collins, I. F., 230, 260, 401, 409
Atkinson, M., 154, 173 Conrad, H., 145, 172
Austen, A., 308 Cook, M.,•138, 172
Avitzur, B., 249, 250, 261, 269,284,294, 305, Cook, P. M., 144, 172
307,309 Crossland, B., 120, 171

Baldwin, W. M., 143, 172, 309 Davis, H. E., 142, 172


Baraya, G. L., 385, 409 Dean, T. A., 149, 173, 273, 305
Barlow, D. A., 339, 346 Dorn, J. E., 87, 89
Bassett, M. B., 81, 89 Drucker, D. C., 5, 10, 81, 89, 105, 118, 349,
Bauscher, J. A., 347 408
Bauschinger, J., 158, 173 Druyanov, A. B., 250, 261
Bedi, D. S., 256, 262 Duby, J., 149, 173
Bell, J. F., 145, 152, 172 Duffin, A. w., 283, 307
Bell, R., 163, 174 Duncan,J. L., 163,165,174,338,346
Bishop, J. F. W., 117, 118, 208, 218, 219, 259,
260,269,275,282,305,306,390,409
Bland, D. R., 113, 118, 249, 257, 261 Eichinger, A., 81, 89, 275, 305
Blazynski, T. Z., 295,309 Eilon, S., 296, 309
Bobrowsky, A., 308 Eisbein, W., 281,306
Boulger, F. W., 308 El-Sebaie, M.G., 341, 342, 346
Bramah, J., 281 Espey, G., 309
Bramley, A. N., 163, 164, 174, 401, 409 Essenburg, F., 85, 89
Brewer, R. C., 249, 261
Bridgman, P. W., 68, 88, 119, 134, 171, 172,
292,308 Fangmeier, E., 281,306
Briggs, G. C., 339, 346 Farren, W. S., 123, 171
Buerk, E., 341, 346 Feldman, H. D., 282, 307
Burgdorf, M., 307 Fiorentino, R. J., 308
Burns, R. S., 156, 173 Firbank, T. C., 250, 261
Fisher, J. C., 151, 173
Flanigan, A. E., 142, 172
Caddell, R. M., 289, 308 Flinn, J. E., 296, 309

*Numbers in italics refer to reference sections at the ends of chapters.


AUTHOR INDEX 417
Fogg, B., 343, 347 Karman, Th. von, 3, 4, 9, 249, 261, 266, 304
Ford, H., 81, 89, 103, 117, 139, 172, 208, 249, Kasuga, Y., 341, 346
250,256,257,259,261,262 Kienzle, 0., 6
Freeman, P., 339, 346 Kobayashi, S., 5, 6, 10, 152, 173, 399,401,409
Frisch, J., 282, 307 Koff, W., 85, 89
Fukui, S., 339, 346 Koiter, W. T., 349, 354, 408, 409
Korber, F., 275, 305
Kuczynski, K., 165, 174
Geckler, J. W., 344, 347 Kudo, H., 5, 10, 227, 250, 260, 262, 282, 306,
Geiringer, H., 4, 9, 187, 259 392,394,398,401,409
Goldsmith, W., 150, 173
Goodier, J. N., 159, 174
Green, A. P., 4, 9, 10, 105, 117, 208, 209, 230, Lancaster, P. R., 250, 261
259, 260, 275, 282, 305, 306, 366, 392, 409 Lankford, W. T., 87, 89, 163, 174, 347
Green, D., 294,308 Larke, E. C., 138, 172, 249, 261
Greenberg, W., 5, 10, 349, 408 Lay, 0. P., 346
Grunzweig, J., 223, 260 Lee, E. H., 222, 231, 260
Gunn, D. A., 308 Leeming, H., 339, 346
Gvozdev, A. A., 5, 10, 349, 408 Lengyel, B., 294, 308
Levy, M., 3, 9, 99, 117
Lianis, G., 81, 89
Haar, A., 3, 9, 266, 304 Lippmann, H., 273, 305
Haddow, J. B., 5, 10, 185, 259, 273, 305 Lloyd, D. H., 347
Haigh, B. P., 74, 88 Lode, W., 4, 9, 81, 88
Hansson, A., 339,346 Loizou, N., 135, 172
Hawkyard, J. B., 269, 305 Longman, I. M., 223,260
Haythornthwaite, R. M., 5, 10, 349, 408 Low, A. H., 308
Hencky, H. Z., 4, 9, 70, 88, 185, 259 Lubahn, J. D., 309, 339, 346
Hessenberg, W. C. F., 256, 262, 328, 329, 346 Ludwik, P., 63, 66, 124, 129, 150, 171
Heyer, R. H., 156, 173 Lyashkov, V. B., 249, 261
Hill, R., 4, 5, 9, 10, 81, 87, 89, 90, 102, 115, 117,
118, 166, 174, 185, 208, 209, 212, 218, 221,
222, 231, 250, 259, 261, 263, 265, 266, 275, Macgregor, C. W., 151, 173
282,296,304,305,306,309,311,339 ,343, MacLellan, G. D. S., 275, 305
346,347,349,382,408 Magnusson, A. W., 143, 172
Hillier, M. J., 256, 262 Mahtab, F. U., 146, 172, 227, 260
Hitchock, J., 250, 258, 261 Malherbe, M. C., de, 165, 174, 383, 409
Hodge, P. G., 159,174, 192,259,282,306,310, Malvern, L. E., 149, 151, 173
311, 345 Manjoine, M. J., 149, 172, 173
Hoffman, 0., 261 Marciniak, Z., 165, 174
Hoffmanner, A. L., 284, 307 Massey, H. F., 385, 409
Hoggart, J. S., 276, 306 McDermott, R. P., 401,409
Hohenemser, K., 105, 117 McKenzie, J., 307
Hopkinson, B., 149, 173 McLean, D., 159, 174
Huber, M. T., 70, 88 McShane, I.E., 230, 260
Hundy, B. B., 105, 117 Melan, E., 112, 118
Mellor, P. B., 5, 10, 149, 163, 165, 172, 174,
283,307,341,342,346
Il'yushin, A. A., 325, 346 Mises, R., von, 3, 9, 69, 88, 99, 117
Inouye, K., 151, 173 Mohr, 0., 30, 36
Moore, G. G., 295, 309
Morgan, R. A. P., 282, 307
Jackson, L. R., 87, 89, 163, 174 Morrison, H. L., 263, 304
Jevons, J.D., 339, 346 Morrison, J. L. M., 105, 118
Jimma, T., 343, 347 Musiol, K., 275, 305
Johnson, R. W., 292, 308
Johnson, W., 5, 10, 146, 149, 163, 165, 172,
173,174,185,227,230,250,259,260 ,262, Nadai, A., 139, 149, 171, 172
269, 276, 282, 305, 306, 338, 343, 346, 347, Naghdi, P. M., 85, 89
359,383,385,401,409 Naylor, H., 276, 305
418 AUTHOR INDEX

Odquist, F. K. G., 4, 9 Snyder, S. 0., 347


Orowan, E., 145, 172, 249, 261 Solkolnikoff, E. S., 115, 118
Osgood, W. R., 131, 172 Solkolnikoff, I. S., 36, liS, 117, 118
Oxley, P. L. B., 186, 259 Sowerby, R., 5, 10, 165, 174, 185, 259, 276,
306,343,347
Stack, E. A., 308
Palmer, W. B., 186, 259 Stockton, F. D., 81, 89
Parker, E. R., 142, 172 Stone, M. D., 250, 261
Parker, J., 81,89 Sturgess, C. E. N., 149, 173
Parkins, R.N., 281, 306 Suiter, J., 276, 306
Pearce, R., 165, 174 Swift, H. W., 295, 296, 309, 328, 339, 343, 346,
Pearson, C. E., 281,306 347
Petch, N.J., 223,260
Phillips, K. A., 149, 150, 151, 172
Pozdeyev, A. A., 249,261 Tanner, R. 1., 385, 409
Prager, W.,5,10, 131,188,192,205,208,259, Tarnovskii, I. Y., 249, 261
282,306,311,345,349,408 Taylor, G. 1., 4, 9, 81, 88, 104, 123, 171
Prandtl, L., 4, 9, 96, 117, 217, 221, 231, 260 Thomas, T. Y., 199, 259
Pugh, H. Ll. D., 105, 117, 293, 307, 308 Thomsen, E. G., 6, 10, 152, 173, 282, 307
Thomson, P. F., 276,306
Tipper, C. F., 143, 172
Quinney, H., 4, 9, 81, 88, 104 Tracy, D. P., 309, 339, 346
Tresca, H., 3, 9, 70, 88
Tselikov, A. 1., 249, 261
Ramberg, W., 131, 172 Tsutsumi, S., 341, 346
Reuss, A., 4, 9, 96, 117 Tucker, G. E. G., 343, 347
Richmond, 0., 263, 304 Tupper, S. J., 222, 231, 260, 275, 305
Ripperberger, E. A., 152, 173
Ros, M., 81, 89
Rowe, G. W., 249, 261 Underwood, L. R., 249,261
Unksov, E. P., 171

Sabroff, A. M., 308


Sachs, G., 4, 9, 242, 260, 261, 275, 276, 281, Venter, R., 165, 174, 383, 409
295,305,306,309,339,346 Voce, E., 131, 172
Saint-Venant, B., de, 3, 9, 96, 117
Samanta, S. K., 149, 173
Schey, J. A., 282, 307 Wallace, J. F., 295, 309, 341, 346
Schmidt, R., 4, 9 Watkins, M. T., 307
Schroeder, W., 269,305 Watts, A. B., 139, 172, 256, 262
Sejournet, J., 282, 307 Weber, E., 295, 309
Sendzimir, M. G., 249, 261 Webster, D. A., 269, 305
Senior, B. W., 344, 345, 347 Westergaard, H. M., 74, 88
Shelton, A., 120, 171 Whiteley, R. L., 173, 341,346
Shepherd, W. M., 105, 118 Whitton, P. W., 256, 257, 262, 288, 307
Shield, R. T., 263, 269, 276, 304, 305 Wilcox, R. J., 256, 262, 288, 307
Siebel, E., 4, 9, 81, 89, 266,275, 281,288,295, Willis, J., 328, 346
305,306,307,309 Wise, D. E., 173
Sims, R. B., 135, 172, 256, 262 Wistreich, J. G., 275, 279, 305, 306
Slater, H. K., 294, 308 Woo, D. M., 332, 346
Slater, R. A. C., 146, 172, 365, 385, 409
Smirnov, V. V., 249, 261 Yang, C. T., 6, 10, 152, 173
Smith, K. F., 87, 89, 163, 174
Smith, 0., 275, 305
Smythe, J. A., 281, 306 Zeitlin, A., 292, 308
Subject Index

Aelotropy, 153 Compression tests, 134


Anisotropy, 152 axisymmetric, 134
control of, 156 plane strain, 138
effects of, 156 Couple equilibrium, 19
flow rule, 114 Cutting, 224
normal, 153
orthotropic, 87
planar, 153 Dead metal zone, 136, 211, 219, 224, 229, 233,
principal axes of, 87 376
yield criterion, 86 Deep drawing of cylindrical cup, 328
Axial symmetry, 263 blank holder, 328
equivalent strain-rate for, 263 drawing ratio, 341
fundamental equations for, 263 earing in, 343
stress equations, 265 flange wrinkling in, 330, 343
stress-strain-rate relations, 265 frictional resistance
velocity equations, 266 at die profile, 335
at flange, 334
ironing in, 330, 339
Bar and wire drawing, 276 mean yield stress, 334
back tension, 276 plastic bending over die profile, 335
drawing stress, 277 progressive stages in, 328
homogeneous,279 punch force, 338
maximum fractional reduction, 279 radial drawing of flange, 332, 337
normal die pressure, 278 redrawing, 341
redundant deformation during, 286 thickness history of blank, 330
Barrelling, 136 unbending at die profile, 338
Bauschinger effect, 68, 85, 87, 90, 103, 158 wall stress, 338
Double subscript notation, 14
Drucker's postulate, 105
Cam plastometer, 144
Cauchy problem, 412
Centred-fan field, 193, 212, 217 Earing, 156, 343
Characteristics of partial differential equa- Extremum principles, 349
tions, 185, 412 lower bound
Cold rolling, 248-259, see also Rolling of estimate, 349
plate and strip theorem, 356
Compression of circular cylinder, see Homo- kinematically admissible velocity field, 358
geneous compression, Upsetting plastic potential, 354
Compression of prismatic block, singular points, 354
between rough platens, 231 surfaces
between smooth platens, 229 of stress discontinuity, 351
420 SUBJECT INDEX

of velocity discontinuity, 352 Invariants


uniqueness theorem, 354 deviator
upper bound strain, 60
estimate, 349 stress, 29
theorem, 358 strain, 56
virtual work equation, 349 stress, 23, 29
Extrusion
backward or inverted, 210
forward or direct, 209 Kronecker delta, 29, 91
relative temperature distribution, 389
temperature distribution in, 385
through square face die, 209, 372 Levy-von Mises equations, 99, 141, 177
through wedge-shaped die, 245, 369 Lines
Extrusion through conical die, 280 of thermal discontinuity, 383
discard length, 289 of velocity discontinuity, 360
frictional resistance at billet -container Lode
interface, 289 strain parameter, 104
maximum extrusion pressure, 290 stress parameter, 82
minimum extrusion pressure, 291 Lower bound theorem, 356
normal die pressure, 285 Liiders hands, 122; 127, 128
open die extrusion, 285
pressure, 284
phases during, 282 Maxiinum plastic work dissipation
pipe, 283 principle of, 115
redundant deformation during, 286 Mohr circle diagrams
incremental strain, 65
plane strain, 179
Flow rule, 100, 112 plane stress, 30, 201, 205
anisotropic, 114 strain-rate, 181
Flow stress, 122 Metal forming processes
dynamic, 147 classification of, 5
quasi-static, 147 Metal processing, 2
Force equilibrium, 16
Forming limits, 7
Optimum die angle, 291

flodographs, 187,212,360,364,368,370,378,
379 Piercing, 227
flomogeneous compression of cylindrical Plane plastic stress, 310
billet, 135, 169 bending of thin circular plate, 321
energy dissipated during, 169 eq liations of, 312
increase in temperature during adiabatic slip line field for uniform stress state, 317
compression, 171 thin plate with circular hole, 318
rigid-perfectly plastic, 170 uniform stress state, 316
strain-hardening, 170 von Mises yield criterion for, 318
flomogeneous deformation, 168 Plane strain compression of prismatic block,
flomogeneous drawing of bar and wire, 279 236
flydrostatic bulge test, 162, 165 deformation energy, 239
flydrostatic extrusion, 292 force exerted by platens, 238
flydrostatic stress, 7, 28, 68, 178 friction hill, 238
mean platen pressure, 238
normal platen pressure distribution, 238
Indentation of semi-infinite medium, 217 plane of no slip, 239
by flat rigid punch, 217, 366 transition between Coulomb friction and
with wedge-shaped indenter, 221 stiction, 241
Instability, plastic with constant frictional stress, 240
biaxial tension, 160 with Coulomb friction, 236
circular sheet metal blank, 165 Plane strain deformation, 175
uniaxial tension, 132, 159 equations for, 176
SUBJECT INDEX 421
Plane strain drawing of sheet metal through Strain, 37-66
wedge-shaped die, 242 compatibility equations, 48
back tension, 242 conventional or engineering, 64, 123
drawing stress, 243 deviator, 59
maximum fractional reduction, 244 dilatational
normal die pressure, 244 conventional, 47, 65
redundant deformation during, 245 natural, 65
Plane strain extrusion through wedge-shaped direct, 40
die, 245 engineering shear, 42
extrusion pressure, 245 fmite, 51
redundant deformation during, 245 free thermal, 92
Plane strain slip line field theory, 182 geometry of, 56
centred-fan field, 193, 212, 217 increment oflogarithmic or natural, 63, 124
characteristic curves or characteristics, 185 infinitesimal, 37
compression of prismatic block, 229 invariants of, 56
coronet, 222 deviator, 60
cutting, 224 logarithmic or natural, 63, 124
extrusion through square face die, 209 maximum shear, 56
Geiringer velocity equations, 186, 199, 216 mean, 59
geometrical properties of slip lines, 189 Mohr circle diagram for incremental, 65
Hencky's first theorem, 190 octahedral direct, 58
Hencky's second theorem, 192 octahedral shear, 58
Hencky stress equations, 184 principal, 54
hodograph, 187,212 principal shear, 56
indentation, 217 pure shear, 45
limit curve, 193 relative displacement tensor, 41
piercing, 227 representative or equivalent, 59, 141, 163
Prager's construction, 204 rotations, 43
requirements for complete solution, 208 simple shear, 45
shear lines or slip lines, 182 spherical, 59
stress Strain-rate, 60
boundary conditions, 199 in uniaxial tension and compression, 126
discontinuities, 195 tensor, 60
singularity, 193, 212 Strain ratio or 'R' value, 153
uniform stress state, 192, 217 average, 154
velocity discontinuities, 198, 212 variation in, 154
Plastic bending Strain tensor
of notched bar, 390 deviator, 59
of thin circular plate, 321 finite, 53
equation of moment equilibrium, 323 pure, 46
limiting lateral pressure, 325 rate, 60
perfectly clamped along outer edge, 327 rotation, 46
rate of curvature, 323 spherical, 59
simply supported at outer edge, 326 Stress, 11-36
Plastic hinge, 390 at a point, 11
Plastic potential, 112 average, 12
cartesian tensor, 21
components of, 13
Rolling of plate and strip, cold-, 248 discontinuities, 195, 351
angle of bite, 258 double subscript notation, 14
determination of coefficient of friction, 257 equivalent for anisotropic material, 115
friction hill, 255 hydrostatic, 7, 28, 68, 178
maximum draft, 259 invariants, 23, 29
neutral plane, 252 maximum shear, 26
normal roll pressure, 252 Mohr circle diagram for plane, 30
roll nominal and true, 123
flattening, 258 octahedral normal, 27
force, 256 octahedralshear,27
torque, 256 principal, 22
Rule of Sarrus, 23, 410 principal shear, 24
422 SUBJECT INDEX

proof, 123 axial force-extension diagram, 120


reduced or deviator, 28, 120 flow stress, 122
representative or equivalent, 28, 94, 141, 163 lower yield stress, 122
states, 24 Luders bands, 122, 127, 128
Stress-strain curves, nominal and true stress, 123
biaxial, 165 proof stress, 123
empirical equations for, 129 ultimate tensile strength, 123
plane strain compression, 140 upper yield stress, 121
showing positive and negative increments Upper bound estimates
of work, 106 anisotropic metals, 382
uniaxial compression, 137 axisymmetric deformation, 398
Stress-strain relations, 90 compression of prismatic block, 363, 365
components of stress increment vector, 108 drawing and extrusion through wedge-
convexity of yield locus, 108 shaped die, 369
deviator stress vector, 102, 115 extrusion through square face die, 372
Drucker's postulate, 105 indentation with flat punch, 366
effect of temperature and strain-rate on, 141 plastic bending
elastic, 91 of rectangular plate, 405
elastic strain energy functions, 93 of thin circular plate, 401
flow rule, 100, 112 Upper bound theorem, 358
flow rule for anisotropic material, 114 applied to plane strain deformation, 359
generalised, 105 Upsetting of circular cylinder, 266
Levy-von Mises, 99, 141, 177 energy dissipation, 269
lode plastic strain parameter, 104 fast compression, 271
plastic potential, 112 force exerted by platens, 269
plastic strain increment vector, 102, 115 friction hill, 268
plastic work, 101, 115 normal platen pressure, 268
Prandtl-Reuss, 96
representative or equivalent plastic strain
increment, 97 Wire drawing, 276
work hardening and strain-hardening hypo- Workpiece, 2
theses, 102 Wrinkling of flange in deep drawing, 7, 343
Stress tensor
cartesian, 21
deviator, 29 Yield criteria, 67-88
spherical, 29 experimental verification of, 79
Strip condition, 414 Haigh-Westergaard stress space, 74
initial yield locus, 84
loading function, 85
Temperature lode stress parameter, 82
adiabatic increase in, 171 pi-plane or synoptic plane, 74
distribution in extrusion, 385 strain-hardening yield function, 85
homologous, 146, 148 stress space representation of, 73
jump, 383 subsequent yield locus, 84
recrystallisation, 143, 146, 148, 151 Tresca
transition, 143 hexagon, 72
velocity modified, 151 yield criterion, 70
Thermal discontinuities, 384 von Mises
Tube drawing and sinking, 294 ellipse, 72
drawing force, 299 yield criterion, 69
floating plug, 295 yield
modified Tresca yield criterion, 298 locus, 72
moving cylindrical mandrel, 303 surface, 71
plug drawing, 295, 301 Yield point, 121
variation of wall thickness, 299 Yield stress
delay, 149
ratio of dynamic to quasi-static, 147
Uniaxial tensile test, 120

Das könnte Ihnen auch gefallen